Next Article in Journal
Modeling and Validation of a LiOH Production Process by Bipolar Membrane Electrodialysis from Concentrated LiCl
Previous Article in Journal
Identification of Suitable Internal Control miRNAs in Bovine Milk Small Extracellular Vesicles for Normalization in Quantitative Real-Time Polymerase Chain Reaction
Previous Article in Special Issue
Investigating the Performance of a Zinc Oxide Impregnated Polyvinyl Alcohol-Based Low-Cost Cation Exchange Membrane in Microbial Fuel Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Bioelectrochemical Systems for Nitrogen and Phosphorus Recovery Using Membranes

Sustainability in Biosystems, Institute of Agrifood Research and Technology (IRTA), Torre Marimon, Caldes de Montbui, E-08140 Barcelona, Spain
*
Authors to whom correspondence should be addressed.
Membranes 2023, 13(2), 186; https://doi.org/10.3390/membranes13020186
Submission received: 1 December 2022 / Revised: 9 January 2023 / Accepted: 26 January 2023 / Published: 2 February 2023
(This article belongs to the Special Issue Membranes for Resource Recovery in Bioelectrochemical Systems)

Abstract

:
Bioelectrochemical systems (BESs) have emerged as a technology that is able to recover resources from different kinds of substrates, especially wastewater. Nutrient recovery, mostly based on membrane reactor configuration, is a clear niche for BES application. The recovery of nitrogen or phosphorus allows for treatment of wastewater while simultaneously collecting a concentrated stream with nutrients that can be reintroduced into the system, becoming a circular economy solution. The aim of this study is to review recent advances in membrane-based BESs for nitrogen and phosphorus recovery and compare the recovery efficiencies and energy requirements of each system. Finally, there is a discussion of the main issues that arise from using membrane-based BESs. The results presented in this review show that it would be beneficial to intensify research on BESs to improve recovery efficiencies at the lowest construction cost in order to take the final step towards scaling up and commercialising this technology.

1. Introduction

Bioelectrochemical systems (BESs) are bioreactors that are equipped with two electrodes, an anode, and a cathode and use exoelectrogenic microorganisms to catalyse oxidation and/or reduction reactions. BESs have become a highly versatile technology that allows wastewater treatment to be combined with the recovery or production of compounds such as nutrients and energy carriers. BESs can be operated as a stand-alone technology or in combination with other technologies, such as anaerobic digestion, in order to improve the recovery of resources and energy within a circular economy approach [1,2,3,4].
BESs can be used to treat a wide range of organic substrates and wastewaters, such as digested sludge and agroindustry, food industry, and urban wastewater [5]. In a changing paradigm in waste and wastewater management technologies from the purpose of waste disposal to the biorefinery concept, the recovery and reuse of nutrients from wastewater is a priority over their disposal through technologies such as the nitrification–denitrification process [6]. For example, fertilisation activities are highly dependent on ammonia and phosphate, and their recovery from waste streams will reduce the demand for phosphate mined from rock or ammonium produced by industrial processes. Ammonium is industrially produced by nitrogen conversion from atmospheric N2 gas through the Haber–Bosch process, which is energy intensive due to the need to supply high temperatures (400–500 °C) and high pressure (100–200 atm) to the reactors. Furthermore, 1.87 tonnes of carbon dioxide (CO2) is generated with every tonne of NH3 produced by the Haber–Bosch process, contributing 1.2% of global anthropogenic CO2 emissions [7]. With regard to phosphate, most of it currently comes from a non-renewable resource, phosphate rock, which has no feasible substitute and can be contaminated by heavy metals. Mining costs increase over time as P rock becomes less available, constraining further mining action. Without proper management, phosphate rock is projected to be depleted in the next 70–140 years [8].
Although BESs can be configured as a single-chambered cell, the most typical setting is a dual-chambered device consisting of an anodic compartment and a cathodic compartment separated by a membrane. This configuration has been widely investigated for nutrient recovery, with different kinds of membranes tested to increase the selectivity of the recovered compound. The most commonly used membranes in the construction of BESs are ion-exchange membranes (IEM), typically to recover ammonium and phosphates for reuse as fertilisers.
BESs have been the subject of various literature reviews focused on materials and their many applications [9,10] along with nutrient removal and recovery [11,12,13], including membrane and membraneless reactors. The most recent reviews focused specifically on the recovery of nitrogen from a sustainability perspective [14,15]. To the authors’ knowledge, no reviews concerning ammonia and phosphate recovery using membrane-based BESs have to date been published. Furthermore, in the last few years, several papers have dealt with nitrogen and phosphate recovery including new advances and focusing both on achieving high efficiency in nutrient recovery and reducing the associated chemical requirements and energy consumption. Thus, an update on new advances in this field is needed.
The main objectives of this review are to highlight the latest advances in nitrogen and phosphorus recovery using different kinds of membrane-BES configurations and compare the recovery efficiency and energy requirements of each system. In addition, the upscaling of BESs, which is gradually being carried out, is reviewed.

2. Methodology

A literature review was undertaken based on a bibliometric analysis of publications in the field of nutrient recovery using membrane-based BESs. The source used to identify these publications was the Web of Science database. In order to include the most recent advances in this field, the research was performed on studies published between 2016 and 2022. The following main keywords were used for the selection: nutrient, ammonia, ammonium, phosphorus, phosphate, bioelectrochemical systems, microbial fuel cell, and microbial electrolysis cell. The obtained papers were analysed to choose only those related to membrane-based reactors. As a result, 49 papers were selected for inclusion in this review.

3. Principles of Nutrient Recovery in BESs Using Membranes

BESs for nutrient recovery generally consist of an anodic compartment and a cathodic compartment divided by a membrane. Exoelectrogenic bacteria grow on the anode, forming a biofilm, and are capable of oxidising organic matter that is introduced to the anode compartment. The electrons and protons produced by the organic matter oxidation at the anode are directed to the cathode compartment by an external electrical circuit and through the internal membrane, respectively, and are combined by a reduction reaction. The circulation of electrons through the external electrical circuit generates electricity (microbial fuel cell, MFC). A charge imbalance is produced in the cell due to electron transfer from the anode to the cathode, which leads to positive charge migration, such as of protons, through the cation-exchange membrane (CEM, the most common IEM for ammonia recovery). However, when complex substrates are introduced to the anode compartment, other cations such as NH4+ will be in higher concentrations than the protons and preferably transfer (migrate) to the cathode compartment against the protons. In a similar way to ammonium, when an anion-exchange membrane (AEM) is used as a separator, phosphate ions can migrate from the cathode to the anode compartment.
Since cation/anion migration through the CEM or AEM is linked to the amount of charge circulating through the external electrical circuit, the transfer rate can be improved by an external power source connection in addition to producing hydrogen gas or other nonspontaneous reactions in the cathode (microbial electrolysis cell, MEC). Ammonia recovery improves with higher current densities; therefore, externally powered MECs perform better than MFCs for the recovery of ions. On the other hand, in the cathode chamber of CEM-equipped MFCs and MECs, a local high pH is created around the cathode surface through proton consumption or release of hydroxyls, which may improve nutrient recovery, i.e., by inducing ammonium conversion to ammonia or struvite crystallisation [16], as discussed in Section 3.1 and Section 3.2.
Membranes are a key element in nutrient recovery in BESs since they allow for the separation of nitrogen and phosphorus from the wastewater matrix and subsequent recovery in a clean solution, in contrast to BESs without a membrane. Several types of IEMs have been applied in the construction of BESs, such as proton-exchange membranes (PEM), CEMs, AEMs, ultrafiltration, microfiltration, and bipolar membranes. Comprehensive reviews of membrane separators in BESs have been published in recent years [17,18]. Membrane-BESs can be configured in different setups using a unique membrane or a combination of different kinds of membranes, achieving different ways of recovering nutrients, as shown in Figure 1.
Although membranes are of great importance in the recovery of nutrients in BESs, reactor performance has been reported to be significantly affected by IEMs since they generally increase the internal electrical resistance and coulombic efficiencies, but also in relation to capital cost, as they represent up to 40% of the total cost of BESs [19]. Nevertheless, membrane-based BESs for nutrient recovery are a subject of great interest, as demonstrated by the 49 papers published since 2016 included in this review related to ammonia recovery (63%) and combined ammonia and phosphate recovery (37%). Although phosphate recovery or remobilisation from iron phosphate in BESs has been reported [20,21], combined nitrogen and phosphorus recovery systems are usually developed.

3.1. Ammonia Recovery by Stripping and Absorption

In CEM-equipped BESs, NH4+ migrates to the cathode chamber through the membrane combined with other cations (Figure 1a). In order to obtain a pure ammonia solution, a subsequent step of stripping and absorption is usually applied. The stripping and absorption process consists of injecting an air flux through the catholyte to drag dissolved gases, in this case ammonia gas. The gaseous stream then comes into contact with an acidic solution, where the ammonia is absorbed, transforming it again into ammonium. In conventional ammonia stripping, the temperature or pH of the substrate needs to be externally increased to favour a shift in the ammonium/ammonia equilibrium towards the second gaseous species. While in BESs, this step is greatly enhanced with no need for an alkali addition or temperature increase, thus reducing reactant costs and energy consumption due to the catholyte spontaneous increase in pH value [14]. Using a NaCl solution in the cathode compartment, the catholyte pH value has been reported to increase from 9.1 to around 10.8–12.1 at the end of the batch [23].

3.2. Nutrient Recovery by Precipitation

Combined nitrogen and phosphorus recovery can generally be achieved by obtaining a concentrate solution or by struvite precipitation. Struvite is a white salt that crystallises with an orthorhombic structure under alkaline conditions when magnesium, ammonium, and phosphate are present in equimolar concentrations, as shown by reaction (1). Struvite is considered to be a slow-release fertiliser.
Mg2+ + NH4+ + HnPO43−n + 6H2O → MgNH4PO4 + 6H2O + nH+
The cost of struvite crystallisation, which in conventional reactors can be high due to pH adjustment using basic chemicals [24] such as Mg(OH)2, NaOH, and Ca(OH)2, can be reduced thanks to catholyte basification in BESs.

4. Nitrogen Recovery Using Membranes in BESs

Recovery of nitrogen in BESs through the use of membranes has been widely studied in recent years with different configurations. As described in Section 3, in most cases, ammonia recovery in BES is based on ammonium migration in combination with other cations through a CEM by two-chambered cells. Other setups have been developed to increase ammonia recovery efficiencies or decrease costs or electrical energy input. Table 1 presents a summary of the main data from the most recently published works. The specific ammonium recovery rate (gN m−2membrane d−1) is used in this paper, normalising data supplied from the different studies as required, in order to facilitate a comparison of performance identified in the various studies.
This section divides nitrogen recovery BESs into three different groups: the most typical two-chambered cells, the hydrophobic membrane-BES configuration, and multiple compartment cells.

4.1. Two-Chambered BESs Equipped with CEM for Ammonia Recovery

One of the possible applications of BESs with two compartments is ammonia recovery from high-strength wastewater (i.e., livestock manure). Great efforts have been made to optimise the operation. The cathode chamber of an MFC continuously fed with pig slurry was coupled to a stripping and absorption unit, achieving a nitrogen flux of 7.2 g N m−2 d−1 using buffer as the catholyte [25]. In the same study, when shifting to MEC mode, the flux of nitrogen increased to 10.3 g N m−2 d−1, reporting a clear improvement when replacing the catholyte buffer with a NaCl solution (25.5 g N m−2 d−1). Furthermore, due to the increase in the pH of the catholyte, ammonia stripping was boosted, increasing the nitrogen recovery efficiency in the absorption column to 94.3% [25]. With the same setup, but using digested pig slurry, similar removal rates were reported [2,23,26,27].
In contrast, the low concentration of readily biodegradable organics can hinder current density production and thus ammonia migration, as reported when digested sludge concentrate was used to operate a dual-chambered MEC for nitrogen recovery. The supplementation of primary sludge fermentation liquor (25 vol%) with concentrate achieved a current density of 6.4 A m−2, compared with that of <0.15 A m−2 obtained using digested sludge alone, with an ammonium removal efficiency of 53% [28]. This improvement was due to the presence of readily biodegradable organic matter in the concentrate. More recently, an excess of 60% of ammonium was recovered from a mixture of anaerobic digester (AD) concentrate and food wastewater (3:1 v/v ratio) using an MEC. Fertilisers formulated with the recovered ammonia solution were used to grow Arabidopsis thaliana as a model plant, confirming that commercial fertilisers can be replaced by the obtained fertiliser supplemented with phosphorus, potassium, and trace elements [29].
A variation in the usual dual-chambered cell equipped with a CEM between the anode and the cathode compartment is the submersible microbial desalination cell developed by Zhang et al. [22]. The aim of this innovative configuration was to recover ammonium and decrease its level in situ in an anaerobic reactor, and it featured an AEM and a CEM outside the anode and cathode chambers, respectively, in contact with the substrate. The authors reported that the external resistance and the initial ammonia concentration influenced ammonia recovery and that ammonia transportation was not negatively affected by the presence of other cations [22].
Dual-chambered MECs with concomitant ammonia recovery and methane production in a biocathode have also been reported, achieving an ammonium flux through the membrane of 14.46 g m−2 d−1, an average removal efficiency of 30% [30], and an ammonium flux through the membrane of 7 g m−2 d−1 [31]. However, methane production in the biocathodes of these MECs was 79 L m−3 d−1 [30] and 92 L m−3 d−1 [31], respectively.
Recently, the typical configuration of a BES coupled to an external unit for stripping was revisited by adopting a stripping system based on an air diffusion cathode in which air is bubbled. This cathode comprised a high-density polyethylene fibre membrane subjected to a hydrophilic treatment and a layer of unidirectional carbon fibres coated with a Pt-free commercial catalyst. Air was forced through the cathode surface, which provided the oxygen needed for the reduction reaction and allowed for ammonia removal from the catholyte into the acid trap. The BES working with synthetic wastewater recovered 45.3% of the initial nitrogen at a current density of 2.5 A m−2. When using real blackwater as the anolyte, the BES achieved a poor performance due to its low content of organic matter and nitrogen [32]. Energy consumption for gas stripping can be also avoided by a passive separation method, facing the cathode directly to the gas phase. This method was applied in a tubular MEC. It was reported that although active aeration achieved a higher ammonia recovery efficiency, the passive separation’s energy consumption was just 1.3 kWh kgN−1, which was significantly lower than the 2.3 kWh kgN−1 with active aeration [33].
One of the main drawbacks of traditional BESs coupled to stripping units for ammonia recovery is the low solubility of oxygen, which leads to a lack of electron acceptors. Zhang et al. [34] proposed the introduction of a highly soluble iron complex (tartrate-EDTA-Fe) as an electron mediator (EM) to the catholyte to improve the transport to the cathode of the oxidising equivalent produced at the stripping unit. This allows a current to be produced, inducing ammonium migration thanks to the oxidation of the EM at the cathode. Air then re-oxidises the EM in the stripping column while simultaneously separating the ammonia. The ammonium recovery rate in the EM system was 3.8 times greater than in the non-EM control. This strategy appears to be especially important when upscaling BESs due to the stacking and modularisation design commonly employed for BESs, since the use of an EM will simplify the system design, avoiding the aeration of each cathode compartment and concentrating it only at the stripping unit [34].
MECs have been also combined with forward osmosis (MEC-FO) to recover nutrients, energy, and water, achieving recovery efficiencies of 99.7% and 79.5% for ammonium nitrogen and phosphorus, respectively [35]. An MEC–FO system was developed to recover ammonium from landfill leachate in an MEC, and the recovered ammonium was used as a draw solute in FO for water recovery. The leachate from the landfill was fed into the MEC anode compartment, and its effluent then flowed into the FO feed side. Using external CO2, an ammonium bicarbonate solution was prepared with the collected ammonia gas that was then used as the FO draw solution. Thanks to the aeration, ammonia gas was stripped out from the catholyte, thus boosting ammonium removal from the anode chamber [36].
Although laboratory-scale BESs have been shown to be a suitable technology for wastewater treatment, especially when applied to urine ammonia recovery, the upscaling of BESs still presents a challenge. However, steps have recently been taken to understand upscaling effects on BESs. San Martin et al. [37] reported that with an increase in the reactor size from 500 mL to 1000 mL, the ammonia recovery efficiency decreased from 47% to 42% without a significant impact on the current density. However, the main bottleneck was the low biodegradability of the wastes [37]. More recently, a 12 L micro-pilot tubular MEC was developed by Cristiani et al. [38] as a post-treatment for an anaerobic digestion process for simultaneous ammonium recovery and biogas upgrading, which achieved a nitrogen removal efficiency of 36% [38]. The largest dual-chambered CEM BES reported in recent years had a volume of 168 L. Three pilot MECs of this dimension were constructed and installed in different municipal wastewater treatment plants (WWTPs) in Switzerland, generating a renewable chemical base and coextracting abundant species, such as ammonium, from wastewater [39].
Table 1 presents a summary of the main data of the most recent studies using dual-chambered BESs, including their operation mode, type of membrane, anolyte and catholyte, and compartment volume.

4.2. Dual-Chambered BESs Combined with Hydrophobic Membrane Modules

The striping and absorption step coupled to BESs in most of the studies referred to in the previous section is highly energy consuming due to continuous air pumping. Hydrophobic membranes (HM) have been tested as an alternative to overcome this issue. Hydrophobic membranes made of gas-permeable materials, such as polyvinylidene fluoride (PVDF), polypropylene (PP), or polytetrafluoroethylene (PTFE), and in the form of tubular, flat, or hollow-fibre membranes, are being coupled to BESs to improve ammonia recovery. Dissolved ammonia gas in a waste stream or in a liquid can pass through the pores of an HM and can be absorbed in an acid solution on the other side, such as sulphuric acid, to form an ammonium salt.
Kuntke et al. [40] reported the recovery of ammonia using an HM module, also called a transmembrane chemisorption (TMCS) module, integrated in the catholyte chamber of a continuously fed MEC. Inside the gas-permeable tubular HM, ammonia was absorbed in sulphuric acid, giving an ammonium sulphate solution. During the 20-day experimental period, the average ammonium removal was 42%, with a maximum removal of 51%. Ammonium was the main contributor to the charge transport through the CEM since on average it represented 70% of the transport efficiency through the CEM, with a maximum of 104% [40]. Different MECs coupled to TMCS units were operated with synthetic and human urine. Each TMCS unit consisted of a PP tubular membrane and was operated in the same way as in the previous study, although in crossflow mode, with the acid on the outside and the catholyte on the inside of the TMCS membrane. The maximum ammonium removal efficiency was 60.9%, which corresponded to a transport rate of 119 gN m−2 d−1 [42]. This is the highest removal rate reported in the studies included in this review, followed by 86.2 gN m−2 d−1 reported by Zhang et al. in a submersed MFC [22].
A TMCS module was also connected to an upscaled MEC (volume of 2.5 L for both the anode and cathode chambers) [43]. The authors reported an unstable ammonia transport through the TMCS module due to both the unstable current density production of the MEC and variations in the conditions of the catholyte (i.e., NH3/NH4+ content and pH), an effect also reported in later research [45]. A high fluctuation in the pH of the catholyte led to unstable ammonia transport through the TMCS membrane. During stable current production, the ammonium transport efficiency through the CEM was 92 ± 25% with a nitrogen recovery of 31 ± 59% [43].
Finally, an electromethanogenic MEC combined with an ammonia recovery system composed of a flat HM was operated with digested pig slurry. The nitrogen transference through the CEM was 8.7 gN m−2 d−1 on average, achieving a removal efficiency of 31% concomitant with methane production [44].

4.3. BESs with Three or More Chambers for Ammonia Recovery

Three-chambered BESs have been tested for ammonia recovery, combining a CEM with an AEM or HM (Table 2). In most cases, ammonia recovery is combined with the production of other value-added products, such as methane.
Xiao et al. [46] proposed a three-chambered MFC to recover not only ammonium from synthetic wastewater, but also nitrate into value-added nitrogenous fertiliser from the anode and cathode chambers, respectively. In the cathode chamber, ferric nitrate was applied as the electron acceptor, which was also used to evaluate the nitrate recovery efficiency. About 47% of the ammonium in the anode chamber and 83% of the nitrate in the cathode chamber were recovered [46].
Zepilli et al. [47] proposed a three-chambered BES configuration to combine ammonium, chemical oxygen demand (COD), and CO2 removal in the same device. In this configuration, an intermediate accumulation chamber was placed between the anode and cathode chambers. A CEM and an AEM were placed on the anode and cathode sides of the accumulation chamber, respectively, to recover a concentrated solution of bicarbonate and ammonium ions concomitant with methane production in the cathode compartment. This allows the recovered concentrated solution of ammonium bicarbonate to be used for agricultural fertilisation or in advanced biotechnology applications such as autotrophic microalgae growth [47]. With the same configuration, the ammonium recovery rate and efficiency were improved based on the ammonium loading rate. The ammonium recovery efficiencies obtained in the study are the highest reported to date, achieving a maximum of 75.5%. The highest recovery rate was 38 gN m−2 d−1) with a 7.4-fold up-concentration of ammonia, reaching 7483 mg L−1 in the concentrate [48].
The MEC with three compartments with an intermediate accumulation chamber was improved by adopting a new two-sided cathode configuration, with two cathode compartments separated from the intermediate anodic compartment by an AEM and a CEM and connected in parallel by a titanium wire. The two-sided cathode configuration MEC allowed for the removal of CO2 through bioelectromethanogenesis in both cathode chambers, while ammonium migration took place only in the cathode compartment equipped with the CEM. Zeppilli et al. [49] developed this configuration to couple CO2 removal from a gas mixture in order to produce CH4 alongside the recovery of ammonium nitrogen. Compared with the three-chambered MEC with an intermediate accumulation chamber assessed earlier by the same authors [47], the two-sided cathode configuration performed better in terms of energy consumption, CO2 removal, and methane production, maintaining similar anodic performances [49]. The two-sided cathode setup was further modified by replacing the CEM with a PEM. This configuration achieved a 61% ammonia removal efficiency when working in galvanostatic mode (300 mA), compared with 47% when working in potentiostatic mode (+0.2 V vs. standard hydrogen electrode, SHE). The catholyte achieved a concentration of 4060 mgN L−1, representing a concentration factor of 4.7 times with respect to the anodic influent [50].
Finally, as in the case of dual-chambered cells, HMs have been assessed as a way of obtaining an ammonia solution free of other cations. Instead of integrating the ammonia recovery system in the catholyte recirculation loop, as described in Section 4.2 and the most commonly used configuration up to now, a three-compartment system MEC can be developed. This last configuration simplifies the operation and construction of the MEC, due to the reduction in the number of vessels, for a future system scale-up. A laboratory-scale MEC with an anode compartment and a cathode compartment, divided by a CEM, and a third compartment placed next to the cathode compartment, separated by a HM, was used for the treatment of organic and nitrogen high-strength wastewater. The flux of ammonium through the CEM was 3.4 gN m−2 h−1. The ammonia recovery through the HM achieved a flux of 1.5 gN m−2 h−1, mainly governed by the pH value and the ammonia concentration of the catholyte [45].
BESs aimed at ammonia recovery have evolved in recent years into more complex designs in order to increase the number of products recovered or to reduce energy and reactant consumption. For example, although the use of HM in combination with BESs has represented savings in energy consumption compared with stripping due to the avoidance of aeration, as described above, an acidic solution is still needed to absorb the ammonia that passes through the membrane, contributing to most of the reagent’s cost. An innovative system called a bioelectrochemical membrane-absorbed ammonia (BEMAA) system was designed by Zhang et al. [51], combining a bipolar membrane (BPM) with an HM in order to avoid this acid consumption. A BPM can be applied to the production of a base and an acid in a BES, reducing operational costs with the simultaneous recovery of ammonia from wastewater. This system comprised five chambers and four membranes: The anode chamber was placed next to the acid production chamber, using a BPM; next to the acid production chamber and separated by an AEM was the desalination chamber, where the ammonia-rich substrate was introduced. Ammonium migrated through a CEM towards the cathode chamber, and finally, an absorption chamber was placed on the other side of the cathode compartment, while a piece of HM was installed between the cathode and the absorption chambers. The base generated in the cathode chamber increased the pH value of the catholyte to boost the transformation of ammonium into ammonia, while the acid simultaneously supplied by the BPM in the acid production chamber was transported to the absorption chamber where ammonia was recovered via membrane absorption. The ammonia removal and recovery efficiencies were 97.3% and 68.1%, respectively [51].
Table 2 presents a summary of the main data of the most recent studies, including the number of compartments, operation mode, type of membrane, anolyte and catholyte, and volume of the compartments.
Table 2. Recent results on nitrogen recovery using BESs with three or more chambers.
Table 2. Recent results on nitrogen recovery using BESs with three or more chambers.
Configuration
(Number of Chambers)
MembraneAnode
Volume (mL)
Cathode Volume
(mL)
Membrane Surface
(cm2)
AnolyteCatholyteMaximum Current Density
(A m−2)
N Recovery Rate
(gNH4+-N m−2 d−1)
Energy Demand
(kWh kgN−1)
Combined withRef.
MFC (3)PEM100100n.r.SyntheticFe(NO3)3 solutionn.r.n.r.n.r.NO3- recovery[46]
MEC (3)CEM
AEM
353535Synthetic 4.838n.r. [48]
MEC (3)CEM
AEM
353535Reject water 4.0236.1–8.2 [48]
MEC (3)CEM60060020Digested pig slurryNaCl solution1.43.4n.r.HM[45]
MEC (3)PEM
AEM
860860289SyntheticSynthetic3.01 c1.3 cn.r.CH4 production[47]
MEC (3)PEM
AEM
860860289SyntheticSynthetic5.02 c2.2 cn.r.CH4 production[49]
MEC (3)CEM
AEM
860860289SyntheticSynthetic6.82 c19.7 c19.66CH4 production[50]
MEC (3)CEM
AEM
860860289SyntheticSynthetic6.92 c18.2 c30.62CH4 production[50]
MEC (3)CEM
AEM
860860289SyntheticSynthetic10.4 c24.8 c27.46CH4 production[50]
Tubular MDC (3)CEM
AEM
2801600n.r.SyntheticDeionised watern.r.n.r.n.r. [24]
MEC (5)BPM, AEM,
CEM, HM
1206016SyntheticSynthetic2.5 c222.91Acid-base production[51]
c Calculated from data reported in the paper.

4.4. Overview of the Relationship between Nitrogen Removal Rates and Current Density

Data about nitrogen removal rates and current densities reported in the papers included in this review are plotted in Figure 2. The studies have been divided into two clusters. In the first group of results, most of the reported current densities are below 1 A m−2, with another important subgroup having current densities in the range of 1–3 A m−2, although with little differences in the nitrogen recovery rate. Fewer studies are placed with current densities above 4 A m−2, achieving nitrogen recovery rates of between 10 and 26 gN m−2 d−1. For the second group, due to the differences in substrate, configuration, reactor size, and operation mode between the studies, no determinant factor for a high ammonia removal rate could be identified. The maximum ammonium transport rate of 119 gN m−2 d−1 [42] was achieved in a study on the load ratio parameter in BESs. The load ratio (calculated as the ratio between the applied current and the nitrogen loading rate) is an important parameter for evaluating the energy input and efficiency of nitrogen removal in processes driven by the current. Operating the system with a higher than optimal load ratio offers no advantage since it consumes more energy and does not increase the removal efficiency [42,52]. Thus, these studies may have had a better load ratio among other possible factors.

5. Concurrent Nitrogen and Phosphorus Recovery Using Membranes in BES

Several studies have dealt with the recovery of ammonium, phosphates, and struvite in BESs [53]. The single-chambered cell configuration was tested [54], and the struvite precipitation was reported mainly on the cathode, which reduces the cathode performance and hinders the recovery of the salt [55]. Consequently, two-chambered and multiple-chambered cells have been developed (Table 3).

5.1. Dual-Chambered Cells for Nitrogen and Phosphorus Recovery

In a dual-chambered MFC, struvite can be obtained through different approaches, achieving a phosphate recovery of between 83% and 95%. A phosphate-rich substrate can be circulated first in the anode compartment for organic matter removal and then in the cathode compartment. Struvite can precipitate in the cathode compartment when dosing an NH4+ and Mg solution due to the increase in pH (pH > 8) near the cathode. Operating the MFC at a high COD and a high aeration flow rate in the cathode compartment increased the current density production and phosphorus recovery [56]. Ammonia required for struvite crystallisation can also be obtained by ammonia migration through the CEM [57,58]. With this configuration, the use of an FO membrane instead of a CEM has been reported to decrease BES performance [59]. An upscaled 10 L dual-chambered MFC stack (volume of each unit of 1.2 L) fed with urine for three months used the pH increase in the cathode compartment to recover phosphate and ammoniacal nitrogen as struvite in an external reactor with an efficiency of 90% and 46%, respectively [60]. Since in most P-rich effluents, such as digestate, there is not enough magnesium to induce struvite crystallisation, requiring the addition of an external source, an innovative bioelectrochemical MgO acidolysis method was developed. MgO was used to inhibit the acidification of the anolyte, while Mg2+ migrated to the cathode chamber for struvite crystallisation to recover P. However, the presence of ammonium in the anode compartment blocked Mg2+ migration, limiting P recovery [61]. Apart from struvite precipitation, nitrogen and phosphorus can be recovered in dual-chambered BESs by other strategies. In dual-chambered cells equipped with a CEM, 80–90% of ammonium recovery from fermented liquor was achieved in the cathode chamber and around 42% of phosphorus in the anode chamber by the formation of polyphosphate granules and polyhydroxyalkanoates (PHAs) in the biomass present in the effluent. Polyphosphate-accumulating organisms removed phosphate and stored it inside cells as polyphosphate, possibly using it for the synthesis of PHAs [62].
Finally, the recovery of nitrogen by dual-chambered CEM-MECs can be coupled with the recovery of water when combined with FO, as described by Zou et al. [35]. First, ammonium migrated from the anode compartment to the cathode compartment. Then, the anolyte was introduced into the feed chamber of the FO for phosphorus concentration and water recovery. After filtering the concentrated feed solution, struvite was recovered. Furthermore, up to 28.7% of the total energy input could be covered by the hydrogen gas generated, giving a specific energy consumption rate of 5.75 kWh kg−1 of struvite [35].

5.2. Three-Chambered Cells for Nitrogen and Phosphorus Recovery

With the combination of an AEM and a CEM, a three-chambered cell can be constructed for nutrient recovery, as described for ammonium in Section 4.3, with a recovery compartment placed between the anode and cathode compartments. These devices are known as a bioelectroconcentration cells (BEC), a resource recovery microbial fuel cell (RRMFC), or a microbial nutrient recovery cell (MNRC). It has been reported that an MNRC achieved nutrient recovery using wastewater chemical energy. NH4+ and PO43− were recovered at 2.6 and 1.0 times the amount, respectively, of their initial concentrations in wastewater. From the recovery solution, more than 62% of NH4+ and 89% of PO43− were crystallised into struvite that could be used as fertiliser [63]. This configuration has been also tested to treat synthetic urine-containing wastewater. In the treatment, the hydrolysis of urea was increased using electrical and microbial processes. The study found that 42% of the total nitrogen and 37% of PO43− were recovered in the middle compartment [64]. A removal efficiency of 59.7% and 42.8% of ammonia and phosphorus, respectively, was reported in a BEC (8.67 kgN m−3 day−1); however, the concentration of ammonium in the cathode compartment was always higher than the values for the anode effluent. It was reported that approximately 10% of all the nitrogen recovered permeated through the AEM, probably due to a high concentration gradient between the feed and the concentrate, a low permselectivity of the AEM, and the migration of uncharged and/or paired N species such as ammonia. When the concentrate was flash-cooled, relatively large crystals (≥500 μm) were obtained that were identified as NH4HCO3 [65]. In a recent study, a three-chambered MEC configuration was used to recover phosphate and ammonium from high-strength nitrogen and organic wastewater (digested livestock manure) in order to precipitate struvite externally. The phosphate and ammonium maximum average removal efficiencies were 36% and 20%, respectively [66].
Monetti et al. described some of the barriers to using domestic low-ionic conductivity wastewater in a BEC, since this substrate produces low current densities (<2 A m−2), thus achieving a low nutrient recovery. Ammonia was up-concentrated 20.1 times. Potassium, sodium, and other micronutrients essential to plant growth, such as Mg, Ca, Fe, S, and Zn, were also detected in the concentrate. However, in this study, phosphorus precipitated as Mg and Ca salts on the membranes and in the concentrate compartment, and they were not recovered in the liquid concentrate [67]. The low suitability of the substrate was also reported by El Qelish and Mahmoud [68], who designed and operated a self-powered MNRC for the recovery of macronutrients from sludge reject water to obtain a liquid concentrate. The MNRC produced a low current density of 0.98 A m−3. The ammonium removal was of 37.1% with a low up-concentration factor of 0.43. When sludge reject water was amended with livestock wastewater (70:30, v:v), increasing the organic loading rate, the current output increased to a maximum of 14.10 A m−3. Ammonium and phosphate up-concentration factors were then 2.19 and 3.41, respectively [68].
BECs have also been combined with FO for nutrient ion recovery from urine. Once NH4+ and PO43– migrated through the CEM and AEM, respectively, and concentrated in the system’s intermediate compartment, they were further enriched by water extraction from the integrated FO process and were finally recovered as struvite. The ammonia and phosphate removal efficiencies of the system were 83.4% and 88.8%, respectively. Analysis of the operating costs showed that the income, including the struvite recovery and electricity generation, could compensate for the operational costs [69].
Not only have struvite or concentrated nitrogen and phosphorus solutions been recovered from BESs, but microalgae biomass has been as well. A photomicrobial nutrient recovery cell (PNRC) was developed by Jiang et al. [70], which used ion migration for their recovery as microalgae biomass. More than 92% of the total phosphorus, nitrate nitrogen, and ammonium nitrogen were recovered from eutrophic water fed to the internal compartment. Although some of the nitrogen was removed by nitrification and denitrification processes, 69.3% of the ammonia removal was reported to be assimilated by algae, while Chloroella vulgaris was responsible for phosphorus removal. Furthermore, C. vulgaris supplied oxygen by photosynthesis in the cathode compartment, allowing for higher current densities in the system and boosting the removal and enrichment of nutrient ions [70].
In order to increase removal efficiencies in MNRC, various strategies have been tested in recent years, such as modified carbon and stainless steel materials for anodes. Carbon-based anodes were reported to achieve higher removal efficiencies for ammonia and phosphorus, reaching 98% and 99%, respectively [71].

5.3. Membrane Stack Configuration Cells for Nitrogen and Phosphorus Recovery

The membrane stack configuration was developed in order to boost recovery efficiencies [72,73]. Chen et al. presented an in situ nutrient recovery and self-sustaining wastewater purification process developed in a 10 L scale reactor called an advanced microbial nutrient recovery cell (AMNRC). Three recovery compartments and two desalination compartments delimited with AEMs and CEMs were inserted between the anode and cathode chambers. PO43– and NH4+ were concentrated in the recovery solution using only the energy contained in the wastewater by 490% and 150%, respectively, and were crystallised out of the circulation system as struvite. NaOH was added to the recovery solution to increase the pH to 8.5–9 in the crystallisation process, and MgCl2·6H2O was supplemented [73]. Another study investigated for the first time how BESs facilitate urea hydrolysis and recover phosphate and nitrogen from diluted source-separated urine. A stacked microbial nutrient recovery cell (SMNRC) was developed with two nutrient recovery compartments and three nutrient removal compartments. The continuous removal of NH4+ from urine in the anode compartment shifted the reaction balance towards continuous hydrolysis. This allowed for the recovery of nutrients from urine with a low energy consumption and no toxicity concerns, with a recovery ratio of 84.6% and 92.9% for nitrogen and phosphorus, respectively [74]. Following this line of urea hydrolysis induction in BESs and NH4+ and PO43− recovery, a later study achieved a 73.1% nitrogen removal and an 86.2% phosphorus removal using three recovery compartments and two dilution compartments, each of which was separated by a membrane. Thanks to the alkaline pH value of the recovery solution (8.1–8.3) and the relatively high concentration of nutrient ions, struvite was recovered [75]. More recently, a novel microbial reverse-electrodialysis electrolysis struvite-precipitation cell (MRESC), containing 10 pairs of desalination cells, operated in fed-batch mode, achieved a phosphate removal efficiency of 72% with a struvite crystallisation rate of 7.62 g m−2 h−1 [76].
A microbial electrochemical nutrient recovery cell (MRC) featuring two desalination compartments was introduced into an anaerobic osmotic membrane bioreactor (AnOMBR), which integrated a FO membrane for physiochemical separation. By using the current density generated in the MRC, phosphate and ammonia were recovered from the bulk solution in the AnOMBR. This integration minimised the diffusion of ammonium through the FO membrane, thus improving the effluent quality and the scaling potential produced by the phosphate. Furthermore, up to 65% of the PO43− and 45% of the NH4+ were recovered from the influent [77].

6. Electricity Consumption of Nutrient Recovery Using Membrane-BES

While most of the recently published studies focusing on ammonia recovery using membrane-BESs include data related to energy consumption estimations (Table 1, Table 2 and Table 3), investigations into the combined recovery of nitrogen and phosphorus are less common.
Generally, the electricity consumption of BESs for nitrogen recovery is compared with the energetic cost of producing fertilisers by the Haber–Bosch process, which accounts for 8.5 kWh kgN−1. However, this viewpoint does not consider other issues from fertiliser production, such as carbon emissions related to fertiliser formulation and distribution, downstream processing, and transport. However, part of the nitrogen produced will end up in wastewater, which will need an energy-consuming treatment. For this reason, recent research is also taking into account the energy needed for the nitrification/denitrification process in a wastewater treatment plant, which accounts for 12.5 kWh kgN−1 [38], or the ammonia oxidation process (Anammox®) that has a consumption of 4.2 kWh kgN−1 [78]. Most of the established recovery technologies, such as air stripping, struvite precipitation, electrodialysis, and adsorption, are still not competitive compared with the Haber–Bosch process that benefits from dimensions of scale [79].
The energy consumption reported in the studies included in this review was in the range of 1.17–2.7 kWh kgN−1, generally considering only the energy supplied for MEC mode operation. The recovery of nitrogen by CEM-MECs combined with FO developed by Zou et al. [35] showed the lowest energy consumption, considering that the hydrogen gas generated could cover up to 28.7% of the total energy input, giving a specific energy consumption rate of 1.17 kWh kgN−1 recovered [35]. When recirculation or aeration for stripping was included, the values for energy consumption increased to 5.7 kWh kgN−1 [80] or more. Qin et al. [36] reported 8.5 kWh kgN−1 in an MEC from which 64.4% of the energy was due to the external power supply, 22.6% was consumed by the recirculation of the anolyte, and the last 13.0% was consumed by aeration in the cathode [36]. The values shown in Table 3 for three-chambered cells are in the range of 20–30 kWh kgN−1, but all the reported values correspond to methane-producing BESs, increasing the output products of the reactor.
Using an electrochemical cell, i.e., a cell with an abiotic anode independent of organic matter, a high nitrogen transport rate at a low energy input was reported. With a current density of 20 A m−2 and an energy demand of 7.25 kWh kgN−1, the nitrogen removal rate from the influent was 151 gN m−2 d−1. The current density of 50 A m−2 produced the maximum nitrogen transport rate (335 gN m−2 d−1) with an energy demand of 15.6 kWh kgN−1 [81]. Although these removal rates are higher than those obtained with BESs due to the higher current densities applied, the energy demand is also higher due to the oxidation of water instead of the oxidation of an organic substrate. However, working in MFC mode reduced the energy consumption to 0.87 kWh kgN−1 due to electricity only being required for aeration since the MFC produced the energy required for ammonia migration [22].
Some authors have evaluated the energy consumption both in galvanostatic and potentiostatic mode operations, since potentiostatic control is the one used in laboratory-scale reactors, especially for anodic biofilm selection, while a galvanostatic control will be a more common electrochemical technique adopted in full-scale processes to maintain the fixed current value. Zeppilli et al. [50] reported that, compared with the potentiostatic operation, the galvanostatic condition increased the current density and, in turn, the recovery of ammonium by 27%. However, due to a decrease in anodic overpotential, the galvanostatic condition increased the energy consumption due to water oxidation in the anode instead of bioelectrochemical COD oxidation [50]. Energy consumption with passive and active aeration has also been compared with reports that an MEC would reduce its electrical consumption with passive aeration to 0.9 kW h m−3 of treated water compared with a consumption of 2.1 kW h m−3 of treated water with active aeration [33]. After comparing different aeration rates and voltages, Qin et al. concluded that there is a compensation between ammonia recovery and energy consumption since it was reported that a high aeration rate in the cathode compartment improved the ammonia recovery with a low energy consumption of 4.9 kWh kgN−1, and ammonia recovery was boosted by a low external resistance due to a higher current production, while a low energy consumption was obtained at mild external voltages (e.g., 0.5 V) [80].
Regarding simultaneous nitrogen and phosphorus recovery BESs, most of them are self-sufficient in regard to energy consumption. An SMDC produced net energy (9.5 kWh kgN−1) from ammonia recovery, assuming a 0.09 kWh kgN−1 electric energy output, a 0.9 kWh kgN−1 consumption for aeration, and a 10.3 kWh kgN−1 energy equivalent for NH3 production using the energy consumption for the Haber–Bosch process [22]. Other studies reported that, compared with conventional technologies, the energy consumption of a BEC is not currently competitive when fed with low-conductivity synthetic urine since a high cell voltage is required, resulting in a relatively low current density production and ammonium recovery rate, with a system consumption of 94.4 kWh kgN−1 recovered [67].
In summary, although the energy consumption is favourable for nitrogen recovery in BES compared with existing technologies, the slow advances in scaling up BESs are hindering their ability to compete with processes already developed on an industrial scale (i.e., Haber–Bosch). An in-depth study of the different parameters involved in electricity consumption in nutrient recovery BESs would be of great importance for upscaling processes since the data supplied from most of the studies included in this review were only based on the voltage applied.

7. Challenges and Future Perspectives

The main challenges of BES application in wastewater treatment have been addressed by different authors, such as upscaling considerations [82] or practical applications [83,84]. Some of the issues are highlighted below.
Although field trials and industrial applications of BESs have been conducted and multiple start-up companies have invested in scaled-up BES demonstrations for the treatment of industrial effluents [85], experiences related to nutrient recovery are still limited. This absence of pilot-scale demonstrations could be attributed to the high cost of membranes and the lack of appropriate electrode modules designed for large-scale applications. Steps are being made towards replacing expensive materials and developing electrodes. The use of graphite-coated stainless steel and carbon felt to fabricate a scalable composite-designed electrode module (total volume of 1 m3) has recently been reported with the aim of integrating mainstream wastewater treatments into BESs [86]. The fabrication of separators, electrodes and, structural frames using biocompatible, low-tech, and fully recyclable materials may help expand their applicability, e.g., rubber or starch-based membranes or electrodes of lignocellulosic biomass pyrolysis charcoal (biochar) as electroactive and conductive materials [87].
However, most of the studies reviewed in this paper are still based on synthetic substrates and highlight the difficulty of achieving high recovery efficiencies with real wastewater due to low conductivity or COD that limit the current density. Since nitrogen and phosphorus recovery through membranes is mainly driven by charge neutrality or migration, boosting the current density is essential if recovery efficiencies are to be improved. Multiple-anode modules can significantly enhance current densities [88,89]. Furthermore, real wastewater’s complex ion composition can compete with ammonia and phosphate recovery or facilitate back diffusion [51]. Wastewater’s ionic composition may affect the transport numbers across the CEM [41]; therefore, it is desirable for the membranes to have a greater ion selectivity. The addition of ionophores to CEMs can selectively transport ammonium through the membrane since they provide a binding site with a suitable size for ammonium complexation by bonding hydrogen with the ionophore components [90].
The suitable operation mode in scaled-up systems may differ from that used in laboratory-scale assays and is another important issue that needs to be addressed. Laboratory-scale operation is generally based on poising the anode to a certain potential (chronoamperometric operation mode). For upscaled systems, a chronopotentiometric method, setting the external circuit current density, might be preferred, avoiding fluctuations in the current density and nitrogen flux. However, the potential of the anode may become less suitable for exoelectrogenic bacteria due to changes produced by the applied voltage, the substrate type, and the concentration or the specific microorganisms developed in the biofilm, for example. In turn, lower COD removal efficiencies might be obtained [50].
Nitrogen and phosphate recovery BES configurations vary widely in their products and underlying processes but coincide in their objective of reusing these nutrients for different purposes. The recovered nutrients can be applied directly as fertilisers [29] since it has been reported that chemical fertilisers can be efficiently replaced by the recovered liquid concentrate at a lower cost and with comparable agricultural productivity [68]. However, other possible applications are emerging. Ammonia recovered from wastewater has been proposed as a nitrogen source for microbial protein production, replacing synthetic nitrogen and contributing to the sustainability of the process [91], or for growing bacteria or algae that could be used for the production of biogas or biofuels [92]. In this sense, more research is needed to fully characterise the concentrated solutions recovered from BESs before their reuse in agricultural soils since livestock manure, urine, and other wastewaters may contain significant concentrations of both inorganic (e.g., heavy metals) and organic contaminants (pesticides, herbicides, antibiotics, etc.) as well as pathogenic microbial strains [87].
Finally, the environmental impact of the recovery of nutrients using BESs needs to be evaluated using life cycle assessment (LCA) methodology. Although LCAs of BESs for wastewater treatment have recently been published [93], no specific LCA for nutrient recovery BESs is available. LCAs could help evaluate the environmental benefits of nutrient recovery in BESs concomitant with other products obtained (such as methane or hydrogen) [94] compared with traditional nutrient production technologies (i.e., the Haber–Bosch process and phosphorus mining).

8. Conclusions

BESs for ammonia and phosphate recovery by membranes offer several advantages over other technologies. BES operation favours ammonia volatilisation due to the pH increase in the cathode compartment, without requiring an increase in temperature or the addition of chemicals, unlike conventional stripping and absorption processes. The combination of different kinds of membranes in the same BES reactor could lead to systems that are capable of generating the base and acid needed for ammonia stripping and absorption, thus avoiding the consumption of reactants. The suitable selection of the ratio between the applied current and the nitrogen loading rate can lead to a high ammonium recovery (up to 119 gN m−2 d−1) with energy consumption in the range of 1.2–2.7 kWh kgN−1 in the future, making it more competitive with the Haber–Bosch process. Membrane-based BESs allow for the simultaneous recovery of ammonia and phosphate, both nutrients of great interest in agricultural systems, with reported removal efficiencies of up to 99%. It would be worthwhile to intensify research into BESs to achieve high recovery efficiencies at the lowest construction cost in order to take the final step towards scaling up and commercialisation. The main drawbacks of scaling up BESs, such as the need for expensive materials to construct the reactors, limitations due to electrode distances in the scaled configurations, or operational conditions, need to be addressed in order to achieve their successful introduction in the market of membrane-based BESs for nutrient recovery.

Author Contributions

Conceptualisation, M.C. and A.B.; methodology, M.C. and A.B.; formal analysis, M.C.; investigation, M.C.; resources, A.B.; writing—original draft preparation, M.C.; writing—review and editing, M.C., V.R. and A.B.; supervision, A.B.; project administration, A.B.; funding acquisition, A.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Spanish Ministry of Economy and Competitiveness (RTA2015-00079-C02-01 and PID2020-118830RR-I00 projects). The support of the CERCA Program and the Consolidated Research Group of Sustainability in Biosystems, (ref. 2021 SGR 01568), both from the Generalitat de Catalunya, is also acknowledged.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AD Anaerobic digester
AEM Anion-exchange membrane
AMNR Advanced microbial nutrient recovery cell
BEC Bioelectroconcentration cell
BES Bioelectrochemical system
BEMAA Bioelectrochemical membrane-absorbed ammonia system
BPM Bipolar membrane
CEM Cation-exchange membrane
COD Chemical oxygen demand
EM Electron mediator
FO Forward osmosis
HM Hydrophobic membranes
IEM Ion-exchange membranes
MEC Microbial electrolysis cell
MFC Microbial fuel cell
MNRC Microbial nutrient recovery cell
MRESC Microbial reverse-electrodialysis electrolysis cell
OsBCRS Forward osmosis integrated bioelectroconcentration and recovery system
PBS Phosphate buffer solution
PEM Proton-exchange membrane
PNRC Photomicrobial nutrient recovery cell
PP Polypropylene
PTFE Polytetrafluoroethylene
PVDF Polyvinylidene fluoride
RRMFC Resource recovery microbial fuel cell
SMNRC Stacked microbial nutrient recovery cell
WW Wastewater
WWTPs Wastewater treatment plant

References

  1. Barbosa, S.G.; Rodrigues, T.; Peixoto, L.; Kuntke, P.; Alves, M.M.; Pereira, M.A.; Ter Heijne, A. Anaerobic biological fermentation of urine as a strategy to enhance the performance of a microbial electrolysis cell (MEC). Renew. Energy 2019, 139, 936–943. [Google Scholar] [CrossRef]
  2. Cerrillo, M.; Viñas, M.; Bonmatí, A. Overcoming organic and nitrogen overload in thermophilic anaerobic digestion of pig slurry by coupling a microbial electrolysis cell. Bioresour. Technol. 2016, 216, 362–372. [Google Scholar] [CrossRef] [PubMed]
  3. Cerrillo, M.; Viñas, M.; Bonmatí, A. Unravelling the active microbial community in a thermophilic anaerobic digester-microbial electrolysis cell coupled system under different conditions. Water Res. 2017, 110, 192–201. [Google Scholar] [CrossRef]
  4. Hassanein, A.; Witarsa, F.; Guo, X.; Yong, L.; Lansing, S.; Qiu, L. Next generation digestion: Complementing anaerobic digestion (AD) with a novel microbial electrolysis cell (MEC) design. Int. J. Hydrogen Energy 2017, 42, 28681–28689. [Google Scholar] [CrossRef]
  5. Kumar, S.S.; Kumar, V.; Malyan, S.K.; Sharma, J.; Mathimani, T.; Maskarenj, M.S.; Ghosh, P.C.; Pugazhendhi, A. Microbial fuel cells (MFCs) for bioelectrochemical treatment of different wastewater streams. Fuel 2019, 254, 115526. [Google Scholar] [CrossRef]
  6. Vanotti, M.B.; Szogi, A.A.; Millner, P.D.; Loughrin, J.H. Development of a second-generation environmentally superior technology for treatment of swine manure in the USA. Bioresour. Technol. 2009, 100, 5406–5416. [Google Scholar] [CrossRef]
  7. Wang, Y.; Meyer, T.J. A Route to Renewable Energy Triggered by the Haber-Bosch Process. Chem 2019, 5, 496–497. [Google Scholar] [CrossRef]
  8. Li, B.; Boiarkina, I.; Young, B.; Yu, W.; Singhal, N. Prediction of Future Phosphate Rock: A Demand Based Model. J. Environ. Inform. 2018, 31, 41–53. [Google Scholar] [CrossRef]
  9. Al-Sahari, M.; Al-Gheethi, A.; Radin Mohamed, R.M.S.; Noman, E.; Naushad, M.; Rizuan, M.B.; Vo, D.-V.N.; Ismail, N. Green approach and strategies for wastewater treatment using bioelectrochemical systems: A critical review of fundamental concepts, applications, mechanism, and future trends. Chemosphere 2021, 285, 131373. [Google Scholar] [CrossRef]
  10. Bajracharya, S.; Sharma, M.; Mohanakrishna, G.; Dominguez Benneton, X.; Strik, D.P.B.T.B.; Sarma, P.M.; Pant, D. An overview on emerging bioelectrochemical systems (BESs): Technology for sustainable electricity, waste remediation, resource recovery, chemical production and beyond. Renew. Energy 2016, 98, 153–170. [Google Scholar] [CrossRef]
  11. Kelly, P.T.; He, Z. Nutrients removal and recovery in bioelectrochemical systems: A review. Bioresour. Technol. 2014, 153, 351–360. [Google Scholar] [CrossRef] [PubMed]
  12. Li, N.; Wan, Y.; Wang, X. Nutrient conversion and recovery from wastewater using electroactive bacteria. Sci. Total Environ. 2020, 706, 135690. [Google Scholar] [CrossRef] [PubMed]
  13. Rodríguez Arredondo, M.; Kuntke, P.; Jeremiasse, A.W.; Sleutels, T.H.J.A.; Buisman, C.J.N.; ter Heijne, A. Bioelectrochemical systems for nitrogen removal and recovery from wastewater. Environ. Sci. Water Res. Technol. 2015, 1, 22–33. [Google Scholar] [CrossRef]
  14. Kuntke, P.; Sleutels, T.H.J.A.; Rodríguez Arredondo, M.; Georg, S.; Barbosa, S.G.; ter Heijne, A.; Hamelers, H.V.M.; Buisman, C.J.N. (Bio)electrochemical ammonia recovery: Progress and perspectives. Appl. Microbiol. Biotechnol. 2018, 102, 3865–3878. [Google Scholar] [CrossRef]
  15. Yang, K.; Qin, M. The Application of Cation Exchange Membranes in Electrochemical Systems for Ammonia Recovery from Wastewater. Membranes 2021, 11, 494. [Google Scholar] [CrossRef] [PubMed]
  16. Zeppilli, M.; Paiano, P.; Torres, C.; Pant, D. A critical evaluation of the pH split and associated effects in bioelectrochemical processes. Chem. Eng. J. 2021, 422, 130155. [Google Scholar] [CrossRef]
  17. Bakonyi, P.; Koók, L.; Kumar, G.; Tóth, G.; Rózsenberszki, T.; Nguyen, D.D.; Chang, S.W.; Zhen, G.; Bélafi-Bakó, K.; Nemestóthy, N. Architectural engineering of bioelectrochemical systems from the perspective of polymeric membrane separators: A comprehensive update on recent progress and future prospects. J. Memb. Sci. 2018, 564, 508–522. [Google Scholar] [CrossRef]
  18. Daud, S.M.; Kim, B.H.; Ghasemi, M.; Daud, W.R.W. Separators used in microbial electrochemical technologies: Current status and future prospects. Bioresour. Technol. 2015, 195, 170–179. [Google Scholar] [CrossRef]
  19. San-Martín, M.I.; Carmona, F.J.; Alonso, R.M.; Prádanos, P.; Morán, A.; Escapa, A. Assessing the ageing process of cation exchange membranes in bioelectrochemical systems. Int. J. Hydrogen Energy 2019, 44, 25287–25296. [Google Scholar] [CrossRef]
  20. Happe, M.; Sugnaux, M.; Cachelin, C.P.; Stauffer, M.; Zufferey, G.; Kahoun, T.; Salamin, P.-A.; Egli, T.; Comninellis, C.; Grogg, A.-F.; et al. Scale-up of phosphate remobilization from sewage sludge in a microbial fuel cell. Bioresour. Technol. 2016, 200, 435–443. [Google Scholar] [CrossRef]
  21. Blatter, M.; Vermeille, M.; Furrer, C.; Pouget, G.; Fischer, F. Mechanisms and Model Process Parameters in Bioelectrochemical Wet Phosphate Recovery from Iron Phosphate Sewage Sludge. ACS Sustain. Chem. Eng. 2019, 7, 5856–5866. [Google Scholar] [CrossRef]
  22. Zhang, Y.; Angelidaki, I. Counteracting ammonia inhibition during anaerobic digestion by recovery using submersible microbial desalination cell. Biotechnol. Bioeng. 2015, 112, 1478–1482. [Google Scholar] [CrossRef] [PubMed]
  23. Cerrillo, M.; Oliveras, J.; Viñas, M.; Bonmatí, A. Comparative assessment of raw and digested pig slurry treatment in bioelectrochemical systems. Bioelectrochemistry 2016, 110, 69–78. [Google Scholar] [CrossRef]
  24. Jing, H.-P.; Li, Y.; Wang, X.; Zhao, J.; Xia, S. Simultaneous recovery of phosphate, ammonium and humic acid from wastewater using a biochar supported Mg(OH) 2 /bentonite composite. Environ. Sci. Water Res. Technol. 2019, 5, 931–943. [Google Scholar] [CrossRef]
  25. Sotres, A.; Cerrillo, M.; Viñas, M.; Bonmatí, A. Nitrogen recovery from pig slurry in a two-chambered bioelectrochemical system. Bioresour. Technol. 2015, 194, 373–382. [Google Scholar] [CrossRef] [PubMed]
  26. Cerrillo, M.; Viñas, M.; Bonmatí, A. Removal of volatile fatty acids and ammonia recovery from unstable anaerobic digesters with a microbial electrolysis cell. Bioresour. Technol. 2016, 219, 348–356. [Google Scholar] [CrossRef]
  27. Cerrillo, M.; Viñas, M.; Bonmatí, A. Microbial fuel cells for polishing effluents of anaerobic digesters under inhibition, due to organic and nitrogen overloads. J. Chem. Technol. Biotechnol. 2017, 92, 2912–2920. [Google Scholar] [CrossRef]
  28. Barua, S.; Zakaria, B.S.; Chung, T.; Hai, F.I.; Haile, T.; Al-Mamun, A.; Dhar, B.R. Microbial electrolysis followed by chemical precipitation for effective nutrients recovery from digested sludge centrate in WWTPs. Chem. Eng. J. 2019, 361, 256–265. [Google Scholar] [CrossRef]
  29. Liu, F.; Worland, A.; Tang, Y.; Moustafa, H.; Hassouna, M.S.E.-D.; He, Z. Microbial electrochemical ammonia recovery from anaerobic digester centrate and subsequent application to fertilize Arabidopsis thaliana. Water Res. 2022, 220, 118667. [Google Scholar] [CrossRef]
  30. Cerrillo, M.; Viñas, M.; Bonmatí, A. Anaerobic digestion and electromethanogenic microbial electrolysis cell integrated system: Increased stability and recovery of ammonia and methane. Renew. Energy 2018, 120, 178–189. [Google Scholar] [CrossRef] [Green Version]
  31. Zeppilli, M.; Pavesi, D.; Gottardo, M.; Micolucci, F.; Villano, M.; Majone, M. Using effluents from two-phase anaerobic digestion to feed a methane-producing microbial electrolysis. Chem. Eng. J. 2017, 328, 428–433. [Google Scholar] [CrossRef]
  32. Losantos, D.; Aliaguilla, M.; Molognoni, D.; González, M.; Bosch-Jimenez, P.; Sanchis, S.; Guisasola, A.; Borràs, E. Development and optimization of a bioelectrochemical system for ammonium recovery from wastewater as fertilizer. Clean. Eng. Technol. 2021, 4, 100142. [Google Scholar] [CrossRef]
  33. Qin, M.; White, C.; Zou, S.; He, Z. Passive separation of recovered ammonia from catholyte for reduced energy consumption in microbial electrolysis cells. Chem. Eng. J. 2018, 334, 2303–2307. [Google Scholar] [CrossRef]
  34. Zhang, D.-X.; Zhai, S.-Y.; Zeng, R.; Liu, C.-Y.; Zhang, B.; Yu, Z.; Yang, L.-H.; Li, X.-Q.; Hou, Y.-N.; Wang, A.-J.; et al. A tartrate-EDTA-Fe complex mediates electron transfer and enhances ammonia recovery in a bioelectrochemical-stripping system. Environ. Sci. Ecotechnology 2022, 11, 100186. [Google Scholar] [CrossRef] [PubMed]
  35. Zou, S.; Qin, M.; Moreau, Y.; He, Z. Nutrient-energy-water recovery from synthetic sidestream centrate using a microbial electrolysis cell - forward osmosis hybrid system. J. Clean. Prod. 2017, 154, 16–25. [Google Scholar] [CrossRef]
  36. Qin, M.; Molitor, H.; Brazil, B.; Novak, J.T.; He, Z. Recovery of nitrogen and water from landfill leachate by a microbial electrolysis cell–forward osmosis system. Bioresour. Technol. 2016, 200, 485–492. [Google Scholar] [CrossRef] [PubMed]
  37. San-Martín, M.I.; Mateos, R.; Escapa, A.; Morán, A. Understanding nitrogen recovery from wastewater with a high nitrogen concentration using microbial electrolysis cells. J. Environ. Sci. Heal. Part A 2019, 54, 472–477. [Google Scholar] [CrossRef]
  38. Cristiani, L.; Zeppilli, M.; Porcu, C.; Majone, M. Mauro Majone Ammonium Recovery and Biogas Upgrading in a Tubular Micro-Pilot Microbial Electrolysis Cell (MEC). Molecules 2020, 25, 2723. [Google Scholar] [CrossRef]
  39. Blatter, M.; Furrer, C.; Cachelin, C.P.; Fischer, F. Phosphorus, chemical base and other renewables from wastewater with three 168-L microbial electrolysis cells and other unit operations. Chem. Eng. J. 2020, 390, 124502. [Google Scholar] [CrossRef]
  40. Kuntke, P.; Zamora, P.; Saakes, M.; Buisman, C.J.N.; Hamelers, H.V.M. Gas-permeable hydrophobic tubular membranes for ammonia recovery in bio-electrochemical systems. Environ. Sci. Water Res. Technol. 2016, 2, 261–265. [Google Scholar] [CrossRef]
  41. Liu, Y.; Qin, M.; Luo, S.; He, Z.; Qiao, R. Understanding ammonium transport in bioelectrochemical systems towards its recovery. Sci. Rep. 2016, 6, 22547. [Google Scholar] [CrossRef]
  42. Rodríguez Arredondo, M.; Kuntke, P.; ter Heijne, A.; Buisman, C.J.N. The concept of load ratio applied to bioelectrochemical systems for ammonia recovery. J. Chem. Technol. Biotechnol. 2019, 94, 2055–2061. [Google Scholar] [CrossRef]
  43. Zamora, P.; Georgieva, T.; Ter Heijne, A.; Sleutels, T.H.J.A.; Jeremiasse, A.W.; Saakes, M.; Buisman, C.J.N.; Kuntke, P. Ammonia recovery from urine in a scaled-up Microbial Electrolysis Cell. J. Power Sources 2017, 356, 491–499. [Google Scholar] [CrossRef]
  44. Cerrillo, M.; Burgos, L.; Bonmatí, A. Biogas upgrading and ammonia recovery from livestock manure digestates in a combined electromethanogenic biocathode—Hydrophobic membrane system. Energies 2021, 14, 503. [Google Scholar] [CrossRef]
  45. Cerrillo, M.; Burgos, L.; Serrano-Finetti, E.; Riau, V.; Noguerol, J.; Bonmatí, A. Hydrophobic membranes for ammonia recovery from digestates in microbial electrolysis cells: Assessment of different configurations. J. Environ. Chem. Eng. 2021, 9, 105289. [Google Scholar] [CrossRef]
  46. Xiao, Y.; Zheng, Y.; Wu, S.; Yang, Z.-H.; Zhao, F. Nitrogen recovery from wastewater using microbial fuel cells. Front. Environ. Sci. Eng. 2016, 10, 185–191. [Google Scholar] [CrossRef]
  47. Zeppilli, M.; Mattia, A.; Villano, M.; Majone, M. Three-chamber Bioelectrochemical System for Biogas Upgrading and Nutrient Recovery. Fuel Cells 2017, 17, 593–600. [Google Scholar] [CrossRef]
  48. Koskue, V.; Rinta-Kanto, J.M.; Freguia, S.; Ledezma, P.; Kokko, M. Optimising nitrogen recovery from reject water in a 3-chamber bioelectroconcentration cell. Sep. Purif. Technol. 2021, 264, 118428. [Google Scholar] [CrossRef]
  49. Zeppilli, M.; Simoni, M.; Paiano, P.; Majone, M. Two-side cathode microbial electrolysis cell for nutrients recovery and biogas upgrading. Chem. Eng. J. 2019, 370, 466–476. [Google Scholar] [CrossRef]
  50. Zeppilli, M.; Cristiani, L.; Dell’Armi, E.; Villano, M. Potentiostatic vs galvanostatic operation of a Microbial Electrolysis Cell for ammonium recovery and biogas upgrading. Biochem. Eng. J. 2021, 167, 107886. [Google Scholar] [CrossRef]
  51. Zhang, Z.; Wang, Z.; Zhang, J.; Deng, R.; Peng, H.; Guo, Y.; Xiang, P.; Xia, S. Ammonia recovery from wastewater using a bioelectrochemical membrane-absorbed ammonia system with authigenic acid and base. J. Clean. Prod. 2021, 296, 126554. [Google Scholar] [CrossRef]
  52. Rodríguez Arredondo, M.; Kuntke, P.; ter Heijne, A.; Hamelers, H.V.M.; Buisman, C.J.N. Load ratio determines the ammonia recovery and energy input of an electrochemical system. Water Res. 2017, 111, 330–337. [Google Scholar] [CrossRef] [PubMed]
  53. Siciliano, A.; Limonti, C.; Curcio, G.M.; Molinari, R. Advances in struvite precipitation technologies for nutrients removal and recovery from aqueous waste and wastewater. Sustainability 2020, 12, 7538. [Google Scholar] [CrossRef]
  54. Yuan, P.; Kim, Y. Increasing phosphorus recovery from dewatering centrate in microbial electrolysis cells. Biotechnol. Biofuels 2017, 10, 70. [Google Scholar] [CrossRef] [PubMed]
  55. Almatouq, A.; Babatunde, A.O. Concurrent phosphorus recovery and energy generation in mediator-less dual chamber microbial fuel cells: Mechanisms and influencing factors. Int. J. Environ. Res. Public Health 2016, 13, 375. [Google Scholar] [CrossRef] [PubMed]
  56. Almatouq, A.; Babatunde, A.O. Identifying optimized conditions for concurrent electricity production and phosphorus recovery in a mediator-less dual chamber microbial fuel cell. Appl. Energy 2018, 230, 122–134. [Google Scholar] [CrossRef]
  57. Ye, Y.; Ngo, H.H.; Guo, W.; Chang, S.W.; Nguyen, D.D.; Liu, Y.; Ni, B.; Zhang, X. Microbial fuel cell for nutrient recovery and electricity generation from municipal wastewater under different ammonium concentrations. Bioresour. Technol. 2019, 292, 121992. [Google Scholar] [CrossRef]
  58. Ye, Y.; Ngo, H.H.; Guo, W.; Chang, S.W.; Nguyen, D.D.; Liu, Y.; Nghiem, L.D.; Zhang, X.; Wang, J. Effect of organic loading rate on the recovery of nutrients and energy in a dual-chamber microbial fuel cell. Bioresour. Technol. 2019, 281, 367–373. [Google Scholar] [CrossRef]
  59. Ye, Y.; Ngo, H.H.; Guo, W.; Liu, Y.; Chang, S.W.; Nguyen, D.D.; Ren, J.; Liu, Y.; Zhang, X. Feasibility study on a double chamber microbial fuel cell for nutrient recovery from municipal wastewater. Chem. Eng. J. 2019, 358, 236–242. [Google Scholar] [CrossRef]
  60. Sharma, P.; Talekar, G.V.; Mutnuri, S. Demonstration of energy and nutrient recovery from urine by field-scale microbial fuel cell system. Process Biochem. 2021, 101, 89–98. [Google Scholar] [CrossRef]
  61. Wang, Z.; Zhang, J.; Guan, X.; She, L.; Xiang, P.; Xia, S.; Zhang, Z. Bioelectrochemical acidolysis of magnesia to induce struvite crystallization for recovering phosphorus from aqueous solution. J. Environ. Sci. 2019, 85, 119–128. [Google Scholar] [CrossRef] [PubMed]
  62. Wu, D.; Sun, F.; Chua, F.J.D.; Lu, D.; Stuckey, D.C.; Zhou, Y. In-situ power generation and nutrients recovery from waste activated sludge – Long-term performance and system optimization. Chem. Eng. J. 2019, 361, 1207–1214. [Google Scholar] [CrossRef]
  63. Sun, D.; Gao, Y.; Hou, D.; Zuo, K.; Chen, X.; Liang, P.; Zhang, X.; Ren, Z.J.; Huang, X. Energy-neutral sustainable nutrient recovery incorporated with the wastewater purification process in an enlarged microbial nutrient recovery cell. J. Power Sources 2018, 384, 160–164. [Google Scholar] [CrossRef]
  64. Lu, S.; Li, H.; Tan, G.; Wen, F.; Flynn, M.T.; Zhu, X. Resource recovery microbial fuel cells for urine-containing wastewater treatment without external energy consumption. Chem. Eng. J. 2019, 373, 1072–1080. [Google Scholar] [CrossRef]
  65. Ledezma, P.; Jermakka, J.; Keller, J.; Freguia, S. Recovering nitrogen as a solid without chemical dosing: Bio-electroconcentration for recovery of nutrients from urine. Environ. Sci. Technol. Lett. 2017, 4, 119–124. [Google Scholar] [CrossRef]
  66. Cerrillo, M.; Burgos, L.; Noguerol, J.; Riau, V.; Bonmatí, A. Ammonium and Phosphate Recovery in a Three Chambered Microbial Electrolysis Cell: Towards Obtaining Struvite from Livestock Manure. Processes 2021, 9, 1916. [Google Scholar] [CrossRef]
  67. Monetti, J.; Ledezma, P.; Virdis, B.; Freguia, S. Nutrient recovery by bio-electroconcentration is limited by wastewater conductivity. ACS Omega 2019, 4, 2152–2159. [Google Scholar] [CrossRef]
  68. El-Qelish, M.; Mahmoud, M. Overcoming organic matter limitation enables high nutrient recovery from sewage sludge reject water in a self-powered microbial nutrient recovery cell. Sci. Total Environ. 2022, 802, 149851. [Google Scholar] [CrossRef]
  69. Jiang, Q.; Liu, J.; Song, X.; Qiu, Y.; Xue, J.; Shao, Y.; Feng, Y. Energy efficient bioelectro-concentration and recovery system of nutrients from human urine by integrating forward osmosis. Resour. Conserv. Recycl. 2022, 181, 106253. [Google Scholar] [CrossRef]
  70. Jiang, Q.; Song, X.; Liu, J.; Shao, Y.; Feng, Y. Enhanced nutrients enrichment and removal from eutrophic water using a self-sustaining in situ photomicrobial nutrients recovery cell (PNRC). Water Res. 2019, 167, 115097. [Google Scholar] [CrossRef]
  71. Shahid, K.; Ramasamy, D.L.; Kaur, P.; Sillanpää, M.; Pihlajamäki, A. Effect of modified anode on bioenergy harvesting and nutrients removal in a microbial nutrient recovery cell. Bioresour. Technol. 2021, 332, 125077. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, X.; Sun, D.; Zhang, X.; Liang, P.; Huang, X. Novel self-driven microbial nutrient recovery cell with simultaneous wastewater purification. Sci. Rep. 2015, 5, 15744. [Google Scholar] [CrossRef] [PubMed]
  73. Chen, X.; Zhou, H.; Zuo, K.; Zhou, Y.; Wang, Q.; Sun, D.; Gao, Y.; Liang, P.; Zhang, X.; Ren, Z.J.; et al. Self-sustaining advanced wastewater purification and simultaneous in situ nutrient recovery in a novel bioelectrochemical system. Chem. Eng. J. 2017, 330, 692–697. [Google Scholar] [CrossRef]
  74. Chen, X.; Gao, Y.; Hou, D.; Ma, H.; Lu, L.; Sun, D.; Zhang, X.; Liang, P.; Huang, X.; Ren, Z.J. The Microbial Electrochemical Current Accelerates Urea Hydrolysis for Recovery of Nutrients from Source-Separated Urine. Environ. Sci. Technol. Lett. 2017, 4, 305–310. [Google Scholar] [CrossRef]
  75. Gao, Y.; Sun, D.; Wang, H.; Lu, L.; Ma, H.; Wang, L.; Ren, Z.J.; Liang, P.; Zhang, X.; Chen, X.; et al. Urine-powered synergy of nutrient recovery and urine purification in a microbial electrochemical system. Environ. Sci. Water Res. Technol. 2018, 4, 1427–1438. [Google Scholar] [CrossRef]
  76. Song, Y.-H.; Hidayat, S.; Effendi, A.J.; Park, J.-Y. Simultaneous hydrogen production and struvite recovery within a microbial reverse-electrodialysis electrolysis cell. J. Ind. Eng. Chem. 2021, 94, 302–308. [Google Scholar] [CrossRef]
  77. Hou, D.; Lu, L.; Sun, D.; Ge, Z.; Huang, X.; Cath, T.Y.; Ren, Z.J. Microbial electrochemical nutrient recovery in anaerobic osmotic membrane bioreactors. Water Res. 2017, 114, 181–188. [Google Scholar] [CrossRef]
  78. Lackner, S.; Gilbert, E.M.; Vlaeminck, S.E.; Joss, A.; Horn, H.; van Loosdrecht, M.C.M. Full-scale partial nitritation/anammox experiences – An application survey. Water Res. 2014, 55, 292–303. [Google Scholar] [CrossRef] [PubMed]
  79. Matassa, S.; Batstone, D.J.; Hülsen, T.; Schnoor, J.; Verstraete, W. Can Direct Conversion of Used Nitrogen to New Feed and Protein Help Feed the World? Environ. Sci. Technol. 2015, 49, 5247–5254. [Google Scholar] [CrossRef]
  80. Qin, M.; Liu, Y.; Luo, S.; Qiao, R.; He, Z. Integrated experimental and modeling evaluation of energy consumption for ammonia recovery in bioelectrochemical systems. Chem. Eng. J. 2017, 327, 924–931. [Google Scholar] [CrossRef]
  81. Kuntke, P.; Rodríguez Arredondo, M.; Widyakristi, L.; ter Heijne, A.; Sleutels, T.H.J.A.; Hamelers, H.V.M.; Buisman, C.J.N. Hydrogen gas recycling for energy efficient ammonia recovery in electrochemical systems. Environ. Sci. Technol. 2017, 51, 3110–3116. [Google Scholar] [CrossRef] [PubMed]
  82. AlSayed, A.; Soliman, M.; Eldyasti, A. Microbial fuel cells for municipal wastewater treatment: From technology fundamentals to full-scale development. Renew. Sustain. Energy Rev. 2020, 134, 110367. [Google Scholar] [CrossRef]
  83. Do, M.H.; Ngo, H.H.; Guo, W.S.; Liu, Y.; Chang, S.W.; Nguyen, D.D.; Nghiem, L.D.; Ni, B.J. Challenges in the application of microbial fuel cells to wastewater treatment and energy production: A mini review. Sci. Total Environ. 2018, 639, 910–920. [Google Scholar] [CrossRef] [PubMed]
  84. Tabassum, N.; Islam, N.; Ahmed, S. Progress in microbial fuel cells for sustainable management of industrial effluents. Process Biochem. 2021, 106, 20–41. [Google Scholar] [CrossRef]
  85. Jadhav, D.A.; Park, S.-G.; Pandit, S.; Yang, E.; Ali Abdelkareem, M.; Jang, J.-K.; Chae, K.-J. Scalability of microbial electrochemical technologies: Applications and challenges. Bioresour. Technol. 2022, 345, 126498. [Google Scholar] [CrossRef]
  86. Wang, A.-J.; Wang, H.-C.; Cheng, H.-Y.; Liang, B.; Liu, W.-Z.; Han, J.-L.; Zhang, B.; Wang, S.-S. Electrochemistry-stimulated environmental bioremediation: Development of applicable modular electrode and system scale-up. Environ. Sci. Ecotechnol. 2020, 3, 100050. [Google Scholar] [CrossRef]
  87. Goglio, A.; Tucci, M.; Rizzi, B.; Colombo, A.; Cristiani, P.; Schievano, A. Microbial recycling cells (MRCs): A new platform of microbial electrochemical technologies based on biocompatible materials, aimed at cycling carbon and nutrients in agro-food systems. Sci. Total Environ. 2019, 649, 1349–1361. [Google Scholar] [CrossRef]
  88. Ren, L.; Ahn, Y.; Hou, H.; Zhang, F.; Logan, B.E. Electrochemical study of multi-electrode microbial fuel cells under fed-batch and continuous flow conditions. J. Power Sources 2014, 257, 454–460. [Google Scholar] [CrossRef]
  89. Dhar, B.R.; Ryu, H.; Santo Domingo, J.W.; Lee, H.-S. Ohmic resistance affects microbial community and electrochemical kinetics in a multi-anode microbial electrochemical cell. J. Power Sources 2016, 331, 315–321. [Google Scholar] [CrossRef]
  90. Casadellà, A.; Schaetzle, O.; Loos, K. Ammonium across a Selective Polymer Inclusion Membrane: Characterization, Transport, and Selectivity. Macromol. Rapid Commun. 2016, 37, 858–864. [Google Scholar] [CrossRef]
  91. Khoshnevisan, B.; He, L.; Xu, M.; Valverde-Pérez, B.; Sillman, J.; Mitraka, G.-C.; Kougias, P.G.; Zhang, Y.; Yan, S.; Ji, L.; et al. From renewable energy to sustainable protein sources: Advancement, challenges, and future roadmaps. Renew. Sustain. Energy Rev. 2022, 157, 112041. [Google Scholar] [CrossRef]
  92. Acién Fernández, F.G.; Gómez-Serrano, C.; Fernández-Sevilla, J.M. Recovery of Nutrients From Wastewaters Using Microalgae. Front. Sustain. Food Syst. 2018, 2. [Google Scholar] [CrossRef]
  93. Zhang, J.; Yuan, H.; Abu-Reesh, I.M.; He, Z.; Yuan, C. Life Cycle Environmental Impact Comparison of Bioelectrochemical Systems for Wastewater Treatment. Procedia CIRP 2019, 80, 382–388. [Google Scholar] [CrossRef]
  94. Chin, M.Y.; Phuang, Z.X.; Woon, K.S.; Hanafiah, M.M.; Zhang, Z.; Liu, X. Life cycle assessment of bioelectrochemical and integrated microbial fuel cell systems for sustainable wastewater treatment and resource recovery. J. Environ. Manag. 2022, 320, 115778. [Google Scholar] [CrossRef]
Figure 1. Examples of three typical configurations of membrane-based BESs. + and − stand for ions charged positively (i.e., NH4+) and negatively (i.e., PO43−), respectively. (a) Dual-chambered BESs, (b) submersed BESs [22], and (c) three-chambered cell with recovery compartment. The main reactions taking place in BESs are shown in (a) by way of example, where ❶ and ❷ stand for MFC and MEC cathodic reactions, respectively.
Figure 1. Examples of three typical configurations of membrane-based BESs. + and − stand for ions charged positively (i.e., NH4+) and negatively (i.e., PO43−), respectively. (a) Dual-chambered BESs, (b) submersed BESs [22], and (c) three-chambered cell with recovery compartment. The main reactions taking place in BESs are shown in (a) by way of example, where ❶ and ❷ stand for MFC and MEC cathodic reactions, respectively.
Membranes 13 00186 g001
Figure 2. Nitrogen recovery rate plotted against current density for the studies included in this review. Results are clustered into two groups: grey dots for the least efficient N recovery studies, and red triangles for the most efficient N recovery studies.
Figure 2. Nitrogen recovery rate plotted against current density for the studies included in this review. Results are clustered into two groups: grey dots for the least efficient N recovery studies, and red triangles for the most efficient N recovery studies.
Membranes 13 00186 g002
Table 1. Recent results on nitrogen recovery using dual-chambered BESs.
Table 1. Recent results on nitrogen recovery using dual-chambered BESs.
Configuration/Operation ModeMembraneAnode
Volume (mL)
Cathode Volume
(mL)
Membrane Surface
(cm2)
AnolyteCatholyteMaximum Current Density
(A m−2)
N Recovery Rate
(gNH4+-N m−2 d−1)
Energy Demand
(kWh kgN−1)
Combined withRef.
MFCCEM500500168Pig slurryPBS0.157.2 n.r.Stripping[25]
MFCCEM500500168Pig slurryNaCl solution0.073.7 n.r.Stripping[25]
MFCCEM14080100SyntheticNaCl solution1.66.81.6Stripping[32]
MFCCEM500500168Digested pig slurryNaCl solution0.228.86n.r. [23]
MFCCEM500500168Digested pig slurryNaCl solution0.411.19n.r. [27]
MFC submersedCEM
AEM
18189SyntheticNaCl solution886.20.86 b [22]
MECCEM500500168Digested pig slurryNaCl solution0.433.73n.r. [26]
MECCEM40014038Reject water Tap water6.4265.8 [28]
MECCEM320 400Diluted urinen.r.1.7n.r.2.48 [40]
MECCEM20018042SyntheticDeionised water1.8910.2n.r. [41]
MECCEM33633696AD concentrate and food WW Tap water2.6 c14 c2.7 [29]
MECPEM250250n.r.Pig slurryPBS0.510.9n.r. [37]
MECPEM500500n.r.Pig slurryPBS0.58.3n.r. [37]
MECCEM500500168Pig slurryPBS4.8 c10.3 n.r.Stripping[25]
MECCEM500500168Pig slurryNaCl solution2.4 c25.5n.r.Stripping[25]
MECCEM500500168Digested pig slurryNaCl solution2.0112.97n.r.Stripping[2]
MECCEM14080100SyntheticNaCl solution2.59.61.6Stripping[32]
MECCEM14080100BlackwaterNaCl solution1.96.51.6Stripping[32]
MECCEM10018042.3SyntheticDeionised water1.67.15.7 aStripping[33]
MECCEM12018096SyntheticNaCl solution1.56.9n.r.Stripping[34]
MECCEM200300100SyntheticNaCl solution10.6119 1.9HM[42]
MECCEM250025005000UrineNaOH solution1.8n.r.1.36HM[43]
MECCEM909072SyntheticNaCl solution0.8 cn.r.1.17Water and P recovery, FO[35]
MECCEM500500168Digested pig slurrySynthetic2.2 c8.64n.r.HM and CH4 production[44]
MECCEM500500168Digested pig slurrySynthetic4.514.46n.r.CH4 production[30]
MECPEM860860289DigestateSynthetic2.1 c7 cn.r.CH4 production[31]
Tubular MECCEM314050602355SyntheticSynthetic0.25 c3.7 c2.3CH4 production[38]
Tubular MECCEM500500n.r.Landfill leachateDeionised water0.72n.r.8.5Stripping
FO
[36]
Tubular MECCEM11591300836 cSyntheticDeionised water0.15 cn.r.1.3Passive NH3 separation[33]
n.r., data not reported; a including aeration and recirculation energy consumption; b only for aeration (MFC produced an electric energy output of 0.09 kWh kgN); c calculated from data reported in the paper; PBS, phosphate buffer solution; HM, hydrophobic membrane; PEM, proton-exchange membrane; CEM, cation-exchange membrane; FO, forward osmosis; WW, wastewater.
Table 3. Recent results on simultaneous nitrogen and phosphorus recovery using membrane-based BESs.
Table 3. Recent results on simultaneous nitrogen and phosphorus recovery using membrane-based BESs.
ConfigurationMembraneAnode
Volume (mL)
Cathode Volume
(mL)
Membrane Surface
(cm2)
SubstrateMaximum Current Density
(A m−2)
Recovered ProductReference
MFCPEM300300n.r.Syntheticn.r.Struvite[56]
MFCCEM305305n.r.Syntheticn.r.Struvite[57]
MFCCEM4530n.r.Synthetic1.17Struvite[61]
MFCCEM2828n.r.Fermented liquorn.r.Ammonia solution and P-rich biomass[62]
MECAEM and CEM60060020Digested pig slurry0.26Nutrient solution[66]
RRMFCAEM and CEM28287.1Synthetic1.9 cNutrient solution[64]
MNRCAEM and CEM11011055Domestic wastewater0.6 cNutrient solution[63]
MNRCAEM and CEM2222n.r.Wastewater6Nutrient solution[71]
MNRCAEM and CEM22022050Sludge reject water with livestock wastewater0.6 cNutrient solution[68]
BECAEM and CEM200200100Synthetic50Nutrient solution[65]
BECAEM and CEM200200100Domestic wastewater2Nutrient solution[67]
PNRCAEM and CEM20207.1 cSynthetic2Microalgal biomass[70]
MEC-FOCEM and FO909072Synthetic Struvite[35]
OsBCRSAEM, CEM and FO10010025Synthetic1.1 cStruvite[69]
MRESCAEM-CEM stack323512Synthetic7.9Struvite[76]
SMNRCAEM and CEM stack21 c7 c7.1cUrinen.r.Nutrient solution[74]
MNRC3 pairs AEM-CEM315105105Urine2 cStruvite[75]
AMNR3 pairs AEM-CEM40004000n.r.Domestic wastewatern.r.Nutrient solution[73]
c Calculated from data reported in the paper; AMNR, advanced microbial nutrient recovery cell; BEC, bioelectroconcentration cell; MNRC, microbial nutrient recovery cell; MRESC microbial reverse-electrodialysis electrolysis cell; OsBCRS, forward osmosis integrated bioelectroconcentration and recovery system; PNRC, photomicrobial nutrient recovery cell; RRMFC, resource recovery microbial fuel cell; SMNRC, stacked microbial nutrient recovery cell.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cerrillo, M.; Riau, V.; Bonmatí, A. Recent Advances in Bioelectrochemical Systems for Nitrogen and Phosphorus Recovery Using Membranes. Membranes 2023, 13, 186. https://doi.org/10.3390/membranes13020186

AMA Style

Cerrillo M, Riau V, Bonmatí A. Recent Advances in Bioelectrochemical Systems for Nitrogen and Phosphorus Recovery Using Membranes. Membranes. 2023; 13(2):186. https://doi.org/10.3390/membranes13020186

Chicago/Turabian Style

Cerrillo, Míriam, Victor Riau, and August Bonmatí. 2023. "Recent Advances in Bioelectrochemical Systems for Nitrogen and Phosphorus Recovery Using Membranes" Membranes 13, no. 2: 186. https://doi.org/10.3390/membranes13020186

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