Next Article in Journal
Development and Application of a New Open-Source Integrated Surface–Subsurface Flow Model in Plain Farmland
Previous Article in Journal
Research on the Flow Characteristics of Power-Law Fluids in Self-Priming Sewage Pumps
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Post-Removal of Phosphorus from Biologically Treated Wastewater and Recovering It as Fertilizer: Pilot-Scale Attempt—Project PhoReSe

by
Kyriaki Kalaitzidou
1,2,*,
Manassis Mitrakas
1 and
Anastasios Zouboulis
3,*
1
Analytical Chemistry Laboratory, Department of Chemical Engineering, Polytechnic Faculty, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
2
Thessaloniki Water Supply & Sewerage Co S.A., 57008 Thessaloniki, Greece
3
Laboratory of Chemical & Environmental Technology, Department of Chemistry, Faculty of Sciences, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
*
Authors to whom correspondence should be addressed.
Water 2024, 16(11), 1527; https://doi.org/10.3390/w16111527
Submission received: 29 April 2024 / Revised: 20 May 2024 / Accepted: 24 May 2024 / Published: 26 May 2024
(This article belongs to the Section Wastewater Treatment and Reuse)

Abstract

:
The major issue of raw materials’ depletion, and more specifically, of phosphorous (an important fertilizer) has currently become an emergent aspect due to expected depletion problems needing immediate handling. This was the reason for the implementation of the PhoReSe project that aimed to remove and recover phosphorus from the secondary (biologically treated) effluent of a municipal wastewater (biological) treatment plant (WWTP “AINEIA”, located near Thessaloniki, N. Greece), treating the wastewaters of the nearby touristic area. Regarding the phosphorous supplementary removal and recovery treatment options, two methods were examined, initially at the laboratory scale (batch experiments), i.e., (1) the adsorption of phosphorous, and (2) the chemical precipitation of phosphorus. Both methods were further applied at the pilot scale by initially performing the adsorption of phosphorous onto the AquAsZero commercial sorbent, which is a mixed manganese iron oxy-hydroxide, followed by the chemical precipitation of phosphorous implemented after the desorption process of the previously saturated adsorbent. The final precipitate of this procedure was examined as an alternative/supplementary fertilizer, this way returning phosphorus into the natural cycle. These experiments, as applied successfully in at the pilot scale, set the basis for larger-scale relevant applications for similar WWTP facilities.

1. Introduction

Even though the Life Cycle Assessment (LCA) method and the concept of the Circular Bio/Economy have dynamically entered industrialized optimization processes, nevertheless, the linear use of materials is still quite widely applied, resulting in the loss or unnecessary disposal of useful raw materials [1]. Apart from sustainability aspects, the increased processing costs for the achievement of even stricter disposal limits, as imposed by respective regulation norms to avoid potential environmental problems, usually require further potential exploitation of used end-of-life materials, such as phosphorus (P), to reduce the respective costs [2]. Phosphorus is among the important nutrients, pointing out the recycled/recovered options during each step of the wastewater treatment process due to its significance for the growth of plants and living organisms, as well as not being considered as a convenient renewable material in nature [3,4]. The crystallization of struvite (i.e., magnesium ammonium phosphate (MAP), MgNH4PO4·6H2O) during certain wastewater treatment processes, in regard to P-rich streams (usually after the anaerobic digestion of sludge), was once considered as a serious operational problem, but nowadays, it is mainly used to retrieve phosphorus from these streams when the existing conditions (i.e., mainly Mg and nutrients’ content, pH, and temperature) allow for its formation, since it can be subsequently effectively applied as a potential fertilizer, therefore enabling the recovery of phosphorus from wastewaters and its environmental reuse [5,6].
Phosphates’ gradual depletion due to the overuse of fertilizers (expected to be from 7.5 to 12.5 Tg/year during 2005–2050) results in a high phosphorous footprint, which can also be related to methane production (being an important part (around 13–20%) of the total greenhouse gases effects) due to the potential creation of algae blooms in water bodies, caused by the resulting eutrophication problem from their respective mismanagement/disposal, and following their death, which can produce methane under anaerobic conditions, hence seriously affecting climate change overall [7,8]. Moreover, the eutrophication problem of water resources is accelerated by anthropogenic activities that include the disposal of municipal (not-sufficiently treated) wastewaters. An increase in the release of nutrients into water bodies, and more specifically, of nitrogen and phosphorus, can result in severe algal blooms, as well as the presence of potentially harmful cyanobacterial species that could, in turn, produce dangerous cyanotoxins [9,10]. As a global issue, eutrophication must be mitigated by implementing efficient treatment strategies that will allow the removal, recovery, and re-entry of phosphorus into its natural bio-geo-environmental cycle. Since there is a tendency to increase the total phosphorus content in raw wastewaters, the target should be to obtain low residual concentration levels (<1 mg/L) in the treated effluents disposed in water bodies; hence, eliminating the undesirable phenomena of eutrophication [11,12]. In the future, if not recycled, phosphorus scarcity will result in the “phosphogeddon” phenomenon, affecting not only climate change, but food production and quality as well [13]. Phosphorus is an essential element for agriculture and cannot be substituted, making the identification of potential strategies to secure sufficient phosphorous supply and avoid the expected deficiency by 2035 very important [14,15]. These strategies must be implemented worldwide towards the exploitation and recovery of phosphorous from several waste streams [16].
Initially, the respective treatment processes regarding phosphorous are aimed at removing it from wastewater to avoid the adverse environmental effects of its disposal. However, nowadays, the problems of phosphorous depletion have led to practices that promote the closing of the P-cycle by implementing processes to recover and recycle it instead [17]. It should also be noted that most current studies focus on the recovery/recycling of P from P-rich streams (e.g., from food production wastes, from municipal wastewater treatment plants (anaerobically treated sewage sludge), and from aqueous sewage liquors and animal wastes etc.) [15,16,17,18,19,20,21,22]. The European Sustainable Phosphorus Platform (ESPP) annually updates the nutrient recovery technologies catalogue, reporting the applied technologies for phosphorus mostly in the forms of struvite, hydroxyapatite, calcium phosphates, and phosphoric acid. Among the 53 respective technologies in the 2023 report, only 1 was reported for the recovery of phosphorus from lower-concentrated steams [23,24]. Bearing in mind the serious economic consequences that will be negatively affected by the “phosphogeddon” depletion effect, but also the respective environmental benefits, the scientific community is oriented to achieve P-recovery not only from P-rich streams (which are mainly in the form of inorganic phosphates/PO43−), but also from lower-concentrated P-streams, such as secondary (treated) wastewater treatment effluents, as well as the meat industry and steelmaking slag, etc. [17,25].
The relevant mitigation techniques may include chemical, physical, and biological processes. P-rich streams in the wastewater treatment process, when the phosphorus recovery is required, usually include urine influent (when separated), secondary wastewater treated effluents, the anaerobic digester supernatant of sewage sludge treatment, and specifically incinerated sewage sludge ash [26]. The wet chemical (hydrometallurgical) method is a simple technique that allows for the recovery of phosphorus mainly from sewage sludge ash or from the anaerobic digester supernatant in the form of high-purity products, such as struvite. However, it presents a relatively high operating cost due to the added chemical reagents and its operational procedures [27]. Sewage sludge ash can be also treated with more complex and higher-energy-consuming thermal methods to recover phosphorus, producing high-purity products and the minimum pollutant content. Diverse technologies, such as pyrolysis and electrochemistry, can be also applied to sewage sludge to recover phosphorus. Even though pyrolysis shows a high recovery rate, its main problems are the potential dioxin and furans gases produced during this process. Electrochemical methods are still under investigation, but mostly at the laboratory scale [28,29].
On the other hand, adsorption is an easy handling process, showing a high selectivity for the removal of phosphorus and being applicable in aqueous phase streams; usually, a proper pre-treatment step for the enhanced separation of suspended solids is required. This method is also highly dependent on the applied sorbent material’s capacity for capturing phosphorus [30,31].
To achieve phosphorus removal and recovery from treated secondary wastewater effluents, in this case, the P-stream that PhoReSe examined, several chemical agents can be added, such as lime, aluminum, or iron salts. However, due to the cost and required quantities of these agents, the process can become viable economically, but only for P-rich streams, while, at the same time, a secondary treatment to eliminate the remaining potential microbial pollutant loads, etc., may be commonly required [32,33]. On the other hand, the application of biological processes cannot handle the phosphorus content constantly and with a high removal efficiency, with around a 70% treatment efficiency for most of relevant cases, also being a highly temperature-dependent process [34,35]. Moreover, the alternative application of micro- or ultra-filtration membrane-specific processes, apart from their rather high implementation costs, can also result in a supernatant that still contains quite high quantities of phosphorus, hence contributing to secondary pollution problems when disposed of, or requiring additional post-treatment [9,36]. Currently, pilot and full-scale applications are also oriented to the combination of adsorption and precipitation processes, such as those examined in the PhoReSe project [31,37].
Since phosphorus scarcity is one of major issues that must be dealt with in the 21st century, the realization of the PhoReSe project was held within the framework of the circular economy and a sustainable development framework, targeting the environmental achievements of closing the phosphorus loop and deficiency and preventing its forecasted elimination as a raw material source. This review paper aims to highlight the successful completion of the PhoReSe project, recovering phosphates from treated wastewaters and presenting selected results to exhibit the potential of its implementation, as supported by previous relevant publications [31,38,39,40,41,42,43]. The importance of the PhoReSe project lies beyond the state of the art of other relevant studies, as it has produced an applicable solution for the recovery of phosphorus from streams of treated municipal wastewater before its disposal in the receiving water bodies, primarily producing an environmentally valuable product that enables the re-entry/reuse of phosphates as a nutrient source, while, at the same time, avoiding the eutrophication phenomena caused by its disposal in rivers, lakes, and even the sea [44].

2. Materials and Methods

2.1. Statistical Variations Determination

For the realization of this research, it is important to examine the stability of wastewater treatment processes and determine the phosphorus concentration that can be potentially recovered. For this reason, the statistical variations of secondary (biologically treated) effluents from a WWTP, handling about 8 × 103 m3/d of influent and consisting of 7 × 103 m3/d of municipal wastewater and 1 × 103 m3/d of domestic septage waste, were examined for the period of 12 months. The mean higher phosphorus concentrations in this effluent were the determinant values for the design of experimental tests [39], because it is crucial for the recovery process to be able to effectively treat the whole range (min.–max.) of respective phosphorus concentrations [45]. The measurement of the statistical variations in the main parameters was conducted using the IBM SPSS statistic software 23 package.

2.2. WWTP Effluent

To achieve the required residual concentrations, lower than the commonly applied legislation limit for PO43− (i.e., 1 mg total P/L, or 3 mg PO43−/L), the continuous recovery of phosphorus disposed with the treated secondary effluents was examined [11]. The municipal WWTP of “AINEIA” treats wastewaters produced from the nearby touristic areas of Thessaloniki, which usually contain (mean) phosphorus concentrations of 4–6 mg P-PO43−/L [31,38,39,40].
The phosphorus adsorption experiments were conducted by using certain hydrous iron oxyhydroxides, either commercially available (i.e., GEH, Bayoxide, and AquAsZero), or produced in the laboratory (schwertmannite and lepidocrocite), as well as the organic resin PuroliteA200 EMBCL supplied by Purolite, Milan, Italy, aiming to examine whether the disposal of phosphorus effluent was achievable below the regulation limit and could be further recovered. Regarding the recovery of phosphorus, the preliminary desorption, in an alkaline environment, of the P-loaded sorbent was investigated, followed by subsequent chemical precipitation, by using certain salts as coagulant agents [31]. The produced precipitated phosphate salts, respectively, were examined further as potential fertilizers by checking their phosphate content and soil bioavailability [38].

2.3. Adsorption Capacity Evaluation

The adsorption experiments evaluated the adsorption capacity by the batch isotherms, thermodynamics, and pH effect fitted in the Langmuir and Freundlich models, in which a quantity of adsorbent was placed in the phosphate solution and examined for its adsorption capacity; also, by using the Rapid Small-Scale Column Tests (RSSCTs), the adsorption capacity was examined in continuous flow. The examined commercially available adsorbents were the following:
  • GEH (supplied by GEH Wasserchemie GmbH & Co. KG, Osnabrück, Germany), consisting mainly form akaganeite.
  • Bayoxide (supplied by Lanxess, Cologne, Germany), consisting mainly of goethite.
  • AquAsZero (supplied by Loufakis Chemicals S.A., Thessaloniki, Greece), consisting mainly of tetravalent manganese feroxyhyte.
  • The organic resin PuroliteA200 EMBCL (supplied by Purolite, Milan, Italy).
The main physico-chemical characteristics of the above-mentioned adsorbents, as previously reported by Kalaitzidou et al., 2016, are presented in Table 1 [31].
The adsorption capacity Q was calculated by the following Equation (1):
Q = C I n i t i a l C R e s i d u a l C A d s o b e n t
where:
Q: adsorption capacity
CInitial: the initial phosphate concentration of the solution used for the adsorption experiment
CResidual: the residual concentration of phosphates after the adsorption experiment
CAdsorbent: the adsorbent concentration in the solution
In addition, two laboratory-synthesized iron oxyhydroxides were examined, prepared under intensive oxidative conditions after the oxidation and precipitation reaction of FeSO4·H2O (supplied by Loufakis Chemicals S.A., Thessaloniki, Greece, with >99.0% purity), producing, as final products, schwertmannite and lepidocrocite, but only in the experiments examining the temperature effect [46].
The relevant thermodynamic adsorption study was conducted in the NSF water matrix, simulating potable water, as shown in Table 1. The water matrix preparation was based on the guidelines provided by the National Sanitation Foundation (NSF) standard and prepared by using chemical-grade reagents, i.e., NaHCO3 (supplied by Merck, St. Louis. MO, USA, with >99.5% purity), NaNO3 (supplied by Merck, with >99.5% purity), NaH2PO4∙H2O (supplied by Sigma-Aldrich, St. Louis. MO, USA, with >98.0% purity), NaF (supplied by Merck, with >99.0% purity), NaSiO3∙5H2O (supplied by Loufakis Chemicals S.A., with >99.0% purity), CaCl2∙2H2O (supplied by Riedel-de Haen, with >99.0 purity), and MgSO4∙7H2O (supplied by Panreac, Applichem, Castellar del Vallés, Barcelona, Spain, with >98.0% purity) [24]. The desired pH values for the experiments were adjusted by the addition of NaOH (supplied by Loufakis Chemicals S.A., Thessaloniki, Greece, as 50% w/w solution), or by HCl addition (supplied by Chemlab NV, Zedelgem, Belgium, as conc. 37% a.r.) and 2 mM of N, N-Bis(2-hydroxyethyl)-2-amino-ethane-sulfonic acid reagent (denoted as BES and supplied by Alfa Aesar, Haverhill, MA, USA, with >99.0% purity), which was used to stabilize the pH at the desired value, as a buffer [40].

2.4. Chemical Precipitation Experiments

The PhoReSe experimental investigation included an examination of the chemical precipitation of phosphorus, with the addition of commercial forms of inorganic salts, i.e., magnesia (MgO), calcium hydroxide (Ca(OH)2), huntite (Ca.Mg3(CO3)4), ferric chloro-sulfate (FeClSO4), and sodium aluminate (NaAlO2), in various water matrixes, i.e., aquatic solutions of potable or distilled water and of biologically treated wastewater (secondary effluent), containing an initial concentration of about 5 mg P-PO43−/L. These salts were examined as Ca2+, Mg2+, Fe3+, and Al3+ sources to form the respective insoluble precipitates of phosphorus [41]. According to the relevant speciation diagram that corresponds to a phosphate concentration of 10 mg/L in the National Sanitation Foundation (NSF) water matrix, which is used as a typical element concentration in natural waters, the respective calcium and magnesium salts can be formed at pH values of >6 at 293 K with the negatively charged phosphate anions [40,47]. The chemical precipitation experiments were conducted in three different matrixes, i.e., distilled, potable, and treated wastewater after the secondary (biological) wastewater treatment, and the main characteristics are shown in Table 2. The phosphate stock solution was prepared by using anhydrous KH2PO4 (supplied by Panreac, Applichem, Castellar del Vallés, Barcelona, Spain, >98.0% purity).
The coagulation/precipitation procedure was investigated initially by batch experiments, adding the respective coagulant doses to the phosphate solutions and examining the optimum phosphate removal, as well as the required dosages and treatment time. The used chemical agents investigated during this project as “coagulants” for the eventual precipitation of PO43− were the following inorganics:
  • magnesia (MgO) (supplied by Grecian Magnesite SA, Yerakini, Greece, with >94.0% purity),
  • calcium hydroxide (Ca(OH)2) (supplied by CaO Hellas, Thessalonki, Greece, with >90% purity),
  • huntite (Ca.Mg3(CO3)4) (supplied by Sibelco Hellas, Thessaloniki, Greece, with >95% purity),
  • ferric chloro-sulfate (FeClSO4) (supplied by Feri-Tri SA, Thessaloniki, Greece, with 12.5% w/w content),
  • sodium aluminate (NaAlO2) (supplied by Loufakis Chemicals S.A., Greece, with content 25% w/w as Al2O3).

2.5. Scaling up to Pilot Scale

The sustainability aspect was the main target of the PhoReSe project, aiming to engage phosphorus from the secondary treated wastewater effluent back into the phosphorus environmental cycle [48]. The promising results obtained from the laboratory-scale experiments led to the design, construction, and operation of a relevant pilot plant, put in operation in the WWTP of AINEIA (see Supplementary Materials) [49]. The pilot plant set-up also included a membrane pre-filtration unit for the efficient removal of residual suspended solids, an adsorption–desorption column, and a final chemical precipitation tank, alongside a full automation panel for remotely controlling the whole recovery process [31,34]. The respective materials and methods are explained in more detail in the previous relevant publications [31,38,39,40].

2.6. Precipitate Evaluation

The resulting chemical precipitates of the phosphorus in the alkaline effluent after the desorption procedure were investigated in terms of phosphorus bioavailability by the extraction of phosphorus with the Mehlich-3 solution (consisting from 0.001 M EDTA, supplied by Merck, with >99.0% purity; 0.015 M NH4F, supplied by Merck, with >98.0% purity; 0.2 M CH3COOH, supplied by Merck, with >99.8% purity; 0.25 M NH4NO3; 0.013 M HNO3, supplied by Merck, as 65.0% a.r.) or the Olsen solution (NaHCO3 0.5 M) for the acidic and alkaline soil samples, respectively. Moreover, the assessment of the soil samples’ phytotoxicity was conducted by using the commercially available toxicity bioassay ‘‘Phytotox kit micro-biotest’’ [38].

3. Results

3.1. Statistical Analysis

During the PhoReSe project, as reported by Raptopoulou et al., 2016, the statistical results showed that notable variations took place regarding the phosphorus content, especially during the summer period in the secondary treated wastewater effluent, indicating the need for additional post-treatment [39].

3.2. Adsorption Capacity

The results evaluation of the adsorbents’ efficiencies, as reported by Kalaitzidou et al., 2016, showed the superiority of iron-based adsorbents, as compared to the lower adsorption/ion exchange capacity of the Purolite A200 EMBCL material [31]. Figure 1 presents the data fittings, applying the Freundlich and Langmuir isotherm models by plotting the adsorption capacity Q of the material versus the residual concentration (CResidual) after the adsorption experiment, and shows the major findings regarding the high iron-based adsorbents’ removal efficiencies. Data for Purolite A200 EMBCL are not included in this plot due to the near-zero adsorption capacity found (0.5 mg PO43−/gads) [31]. The obtained data fitted the Freundlich model better, due to the presence of heterogeneous adsorption sites, while it is obvious that the AquAsZero adsorbent material presented the optimum adsorption capacity (Q), with Qmax 60 mg PO43−/gads, whereas Q3, which was calculated to comply with the regulation limit of 3 mg PO43−/L, was found to be 57.4 mg PO43−/gads. It is remarkable that the adsorption was favored by the co-existing ions, showing a better adsorption capacity when the adsorbents dispersed in the NSF water matrix. The amorphous FeOOH and AquAsZero presented a lower PZC and higher Ctotal (mmol [OH]/g), as presented in Table 1, and showed a superior exchange capacity compared to the crystalline adsorbents GEH and Bayoxide. The adsorption was favored by the charge on the adsorbent surface and the SO42− content (crystalline or adsorbed), as reported by Kalaitzidou et al., 2020 [31,40]. The higher presence of crystalline SO42− existing in FeOOH, which was 5.4 wt% as compared to the respective value of AquAsZero 1.5 wt%, was the main factor that hindered adsorption when there were co-existing ions. For this reason, AquAsZero presented the optimum sorption results [40].
For a better evaluation of AquAsZero, the effects of different pH values were also examined, i.e., pH 6, 7, and 8, with the optimum adsorption results shown for the lower pH value. Moreover, the effect of temperature on the adsorption of phosphate onto the iron-based adsorbents was thoroughly examined, and an exothermic process was reported when the adsorption took place in the distilled water matrix, while an endothermic process was found for almost all examined iron-based adsorbents, except of Bayoxide, when the matrix was the NSF water. The difference in thermodynamics can be attributed to the co-existing ions in the NSF water matrix that favored the adsorption of phosphates onto the iron-based adsorbents. Higher temperatures seemed to favor phosphate removal by adsorption onto iron oxyhydroxides, with 90% removal occurring after 1 h of contact time. The physisorption mechanism, as previously reported by Kalaitzidou et al., 2016 and 2022, resulted in the efficient/easy desorption of the adsorbed phosphates onto AquAsZero, due to the weak forces that were affected by the alkaline environment that shifted the equilibrium to finally desorb the adsorbed phosphates [31,40].
The removal/separation of phosphates by adsorption onto the iron-based adsorbents was also investigated with the continuous running of Rapid Small-Scale Column Tests (RSSCTs), which can more efficiently simulate full-scale applications than the previous batch tests, by using as a matrix the secondary treated effluent from the WWTP “AINEIA”, which, according to the statistical variation results of Table 1, can contain about 12–17 mg of PO43−/L. The breakthrough curves for the first cycle of adsorption resulted in 34, 31, 19, and 15 mg PO43−/gads for the materials AquaAsZero, FeOOH, Bayoxide, and GEH, respectively, as calculated for the regulation discharge limit of 3 mg PO43−/L. As expected, the different chemical contents of this aqueous matrix negatively affected the adsorption procedure [31].

3.3. Chemical Precipitation

The examined phosphate concentration for the preliminary chemical precipitation experiments was approximately 10 mg PO43−/L. The chemical precipitation of phosphorus, due to the addition of magnesia (MgO), calcium hydroxide (Ca(OH)2), or huntite (Ca.Mg3(CO3)4), requires an alkaline environment; hence, a pH greater than 10 was applied to achieve the required efficient capture/removal of phosphorus, and the residual concentration had to be lower than the set regulation limit of 3 mg/L, by using all examined water matrixes, i.e., potable or distilled water and wastewater (secondary effluent). For the same matrix, the addition of 5 mg Al3+/L in the pH range of 6–8 could achieve removals lower than the required regulation limit of 3 mg PO43−/L, which, however, increased significantly for even higher pH values due to the corresponding increase in aluminum solubility. Moreover, this also occurred for the addition of 7.5 mg Fe3+/L and the reaction ratios of 0.5:1 for Al3+:PO43− and 0.75:1 for Fe3+:PO43−, as estimated for the optimum removal at pH 7 (Figure 2). Additionally, the temperature also affected the PO43− removal, as shown in Figure 3, with the temperature increase slightly favoring removal for the cases of Ca(OH)2 or Ca.Mg3(CO3)4 additions, while slightly decreasing removal for the rest of theexamined chemical agents. In addition, the formed precipitates contained almost 50.1% PO43− for the case of the Ca(OH)2 addition, whereas only 1.2% PO43− for the case of Ca.Mg3(CO3)4 and 19.0% PO43− for MgO, 8.7% PO43− for NaAlO2, and 27.3% PO43− for FeClSO4, when the experimental matrix was distilled water [41].

3.4. Pilot-Plant Operation

Similar to the RSSCTs results, an adsorption capacity of 32 mg PO43−/gads, as calculated for the regulation discharge limit of 3 mg PO43−/L, was reported for the following pilot-scale implementation (Figure 4) by using AquAsZero as the adsorbent medium. The pilot-plant successfully processed the secondary treated effluent from the WWTP “AINEIA” both in the adsorption process with an Empty Bed Velocity (EBV) of 10 m/h, as well as in the desorption mode with an EBV of 5–20 m/h (seven adsorption and six desorption cycles for the pilot plant were performed), where continuous adsorption–desorption cycles with an approximately 20% loss of adsorption sites (and efficiency) between the first and second cycle and 10% more loss between the subsequent adsorption cycles were noticed. This capacity loss can be attributed to the phosphates strongly bound onto the iron oxy hydroxides’ adsorption sites that could not be easily desorbed during the subsequent treatment cycle [31,38,42,48]. More specifically, even though during the RSSCTs experiments, the desorption succeeded by applying the NaOH solution at the pH of 12.5 ± 0.2 achieved in approximately 3 h to complete the adsorbents’ regeneration, permitting the reuse of the adsorbent for the pilot-plant operation, the requirement of a regeneration optimization process was reflected in the non-expected reduction in the adsorption capacity during the subsequent operation cycles, as shown in Figure 5. An even higher pH value (12.8 ± 0.2) was applied in the pilot plant to optimize the regeneration process at the fourth regeneration cycle, favoring the adsorption capacity after the fifth operation cycle. Besides the loss of adsorption capacity, this regeneration process is vital for the economic viability of the overall recovery process [43].

3.5. Fertilizer

The alkaline solution from the desorption process was collected in the sedimentation tank, as shown in Figure 4, where, according to the preliminary chemical precipitation experiments, the contained phosphates can be best precipitated and separated with the presence of Ca2+ or Mg2+ addition of salts. Since the precipitate was intended to be finally used as a fertilizer, the maximum PO43− content in the precipitate was desired; therefore, Ca2+ was selected for this process. The non-toxic precipitate, as collected from the pilot-plant application (Figure 6) for the six adsorption–desorption cycles, was approximately 5 kg and presented a high phosphate content (35.4 wt.%) that could be easily subsequently applied as a phosphate-based fertilizer, mostly in acidic soils, where the phosphate release is expected to be better. At the same time, no adverse effects regarding the plants’ growth may be caused, as reported during the phytotoxicity investigation [31,38].

3.6. PhoReSe Significance

Comparing PhoReSe with the relevant REM NUT® process, which also uses adsorption onto the Cu-form resin Dowex M-4195, a rather expensive adsorbent, and uses the secondary effluent from Massafra (S.E. Italy) Wastewater Treatment Plant as a matrix in the respective pilot plant, it becomes obvious that the AquAsZero adsorbent can treat greater volumes of the secondary wastewater, noting that both comparatively examined effluents contain similar concentration of phosphates; hence, PhoReSe presents, in this way, a lower cost for the whole treatment process [50]. Like the PhoReSe project, the secondary wastewater effluent from another municipal sewage treatment plant was treated at the pilot scale by applying phosphate adsorption onto the ZnFeZ material, which also subsequently subjected to desorption and finally, phosphorus precipitated as struvite. Even though the obtained results in this case study seem quite promising, nevertheless, the operational cost is elevated due to the additional magnetic separator required, whereas that process is operated in batch mode (another disadvantage) [51]. Additionally, a similar adsorption process was reported using the HAIX-Layne RT (HAIX) resin for the phosphate anion exchange from the secondary treated wastewater, with a removal capacity of 4.1 mg PO43−/gresin, which is significantly lower than the capacity found for AquAsZero (32 mg PO43−/gads) [52].
The phosphorus content in the municipal waste streams corresponds to the largest percentage (approximately 23.6%) of the whole lost phosphorus from the respective cycle of use, being, at the same time, the most promising recovery option from secondary sources [53]. Effective circulation from the wastewater streams to finally reach (and fertilize) the soil is of high importance, since phosphorus deposits are being quickly depleted worldwide [54,55]. Therefore, sustainable strategies based on the best available technique should be applied to recover the disposed phosphorus from secondary wastewater treated effluents as an added-value material (fertilizer) that will release phosphorus again in the agricultural soil, partially closing, in this way, the respective material balance and reducing the actual phosphorus needs from virgin (mineral) sources [56].
Future work should be oriented to the optimization of fertilizer production by reducing the CaCO3 content in the precipitate. Moreover, a combination of proper materials, such as the serpentinized solid wastes from Grecian Magnesite SA mining operations, could be also used as Mg2+ sources, recycling two different types of wastes and belonging in the general context of sustainability [57]. A supplementary addition that can be also made in the pilot set-up is that of a second adsorption column operating in parallel to the first one, treating the effluent continuously, even when one of these columns is under the regeneration mode.

4. Conclusions

Phosphorus, due to depletion problems, is considered to be a finite resource, alongside the ecological vulnerability of receiving water bodies towards the presence of excess phosphorus, causing eutrophication problems. This has led to the investigation of post-treatment for the secondary effluents from municipal WWTPs, considering both the phosphorus removal and recovery options. Initially, the statistical variations in phosphorus content in an operation year were determined as 12–17 mg PO43−/L. During the performed (preliminary) chemical precipitation experiments, it was noticeable that the removal of phosphorus could be enhanced by the co-presence in the water to be treated with wastewater common ions; the precipitation process can also be positively influenced (after the desorption/regeneration process) by an increase in alkalinity, rather than by the excess addition of chemical reagents. Calcium and magnesium compounds can effectively remove phosphates when the solution pH value is greater than 9, whereas iron and aluminum salts can effectively remove phosphorus in the pH range of 6–9.
Summarizing, the PhoReSe project achieved effective phosphorus recovery from low-concentrated treated wastewater effluent (in the range of few mg/L) in the form of a fertilizer that can re-introduce phosphorus into agricultural cropland. It is of high importance that the circular economy loop of phosphorus can be (partially) closed, a fact that presents important benefits, such as an increased availability of phosphorus-based fertilizers that, nowadays, face severe scarcity issues, also reducing respected high fertilizers’ cost, as well as the potential to prevent eutrophication problems, since the load of phosphorus in the disposal of post-treated effluents can be significantly lower.
The adsorption evaluation during laboratory-scale experiments pointed out that the AquAsZero material presents the optimum adsorption capacity of 34 mg PO43−/L, an efficiency also verified during the subsequent pilot-scale experiments (32 mg PO43−/L), as calculated for the current disposal regulation limit of 3 mg PO43−L; hence, this indicated the proper design of the examined process and implied its suitability for full-scale application at the operating pH of around 7. Both the (small) columns (at the laboratory scale) and the larger ones at the pilot plant could be effectively regenerated by using the NaOH solution at pH > 12.5, resulting in an enriched PO43− solution that could be further treated by adding Ca2+ as the chemical agent to precipitate phosphorus, producing the respective fertilizers that could contribute to the sustainable/proper management of phosphorous. As presented, the PhoReSe project showed a high removal capacity (32 mg PO43−/gads) and recovery efficiency (>95%), resulting in the superiority of the examined treatment process in comparison with other (relevant) pilot-scale plants previously presented in the literature. The recovered phosphorus can effectively re-enter the phosphorus cycle as a fertilizer, which is produced by using the desorption effluent, adding Ca(OH)2, which leads to the phosphate fertilizer. The significance of this project is the potential of its implementation for treated secondary wastewater effluents from municipal WWTPs, a fact that is of great environmental importance, as it complies with the current regulation limits and is considered as economically feasible, unlike most other implemented technologies. The realization of this project is in accordance also with green chemistry and engineering principles, as well as with sustainable environmental management techniques, by using and producing materials that are environmentally friendly.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/w16111527/s1. Refs. [58,59] are cited in Supplementary Materials.

Author Contributions

Conceptualization, A.Z.; methodology, M.M.; software, K.K.; validation, K.K., A.Z. and M.M.; formal analysis, K.K.; investigation, K.K.; resources, A.Z.; data curation, K.K.; writing—original draft preparation, K.K.; writing—review and editing, A.Z.; visualization, M.M.; supervision, A.Z. and M.M.; project administration, A.Z. and M.M.; funding acquisition, A.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was co-financed by the European Union and the Greek State Program PAVET, Project “PhoReSe—Recovery of Phosphorus from the Secondary Effluent of Municipal Wastewater Treatment”. Funding number: 1525-ΒΕΤ-2013.

Data Availability Statement

No new data were created.

Acknowledgments

During this research two departments from the Aristotle University of Thessaloniki participated, i.e., the departments of Chemistry and of Chemical Engineering, collaborating also with “Aktor” S.A. company, whereas this project additionally approved and supported by the Thessaloniki’s Water Supply and Sewerage Co. (EYATh S.A.)—Department of Plants’ Operation, Maintenance & Environmental Monitoring, which is gratefully appreciated.

Conflicts of Interest

Author Kyriaki Kalaitzidou was employed by the company Thessaloniki Water Supply & Sewerage Co S.A. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Johansson, K.; Perzon, M.; Fröling, M.; Mossakowska, A.; Svanström, M. Sewage sludge handling with phosphorus utilization—Life cycle assessment of four alternatives. J. Clean. Prod. 2008, 16, 135–151. [Google Scholar] [CrossRef]
  2. Pradel, M.; Aissani, L. Environmental impacts of phosphorus recovery from a “product” Life Cycle Assessment perspective: Allocating burdens of wastewater treatment in the production of sludge-based phosphate fertilizers. Sci. Total Environ. 2019, 656, 55–69. [Google Scholar] [CrossRef] [PubMed]
  3. Duan, M.; O’Dwyer, E.; Stuckey, D.C.; Guo, M. Wastewater To Resource: Design of a Sustainable Phosphorus Recovery System. Chem. Open 2019, 8, 1109–1120. [Google Scholar] [CrossRef]
  4. Linderholm, K.; Tillman, A.M.; Mattsson, J.E. Life cycle assessment of phosphorus alternatives for Swedish agriculture. Resour. Conserv. Recycl. 2012, 66, 27–39. [Google Scholar] [CrossRef]
  5. Yan, H.; Shih, K. Effects of calcium and ferric ions on struvite precipitation: A new assessment based on quantitative X-ray diffraction analysis. Water Res. 2016, 95, 310–318. [Google Scholar] [CrossRef] [PubMed]
  6. Kabdaşlı, I.; Kuşçuoğlu, S.; Tünay, O.; Siciliano, A. Assessment of K-Struvite Precipitation as a Means of Nutrient Recovery from Source Separated Human Urine. Sustainability 2022, 14, 1082. [Google Scholar] [CrossRef]
  7. Nijman, T.P.A.; Lemmens, M.; Lurling, M.; Kosten, S.; Welte, C.; Veraart, A.J. Phosphorus control and dredging decrease methane emissions from shallow lakes. Sci. Total Environ. 2022, 847, 157584. [Google Scholar] [CrossRef] [PubMed]
  8. Saraiva, R.; Ferreira, Q.; Rodrigues, G.C.; Oliveira, M. Nanofertilizer Use for Adaptation and Mitigation of the Agriculture/Climate Change Dichotomy Effects. Climate 2023, 11, 129. [Google Scholar] [CrossRef]
  9. Akinnaw, O.S. Eutrophication: Causes, consequences, physical, chemical and biological techniques for mitigation strategies. Environ. Chall. 2023, 12, 100733. [Google Scholar] [CrossRef]
  10. Sahu, A.K.; Mir, S.; Nayak, B.; Baitharu, I. Chapter 13—Sustainable management of eutrophication and problems associated with the algal toxin in ponds and lakes of rural areas. In Water Resources Management for Rural Development; Elsevier: Amsterdam, The Netherlands, 2024; pp. 155–170. [Google Scholar] [CrossRef]
  11. EEC Council. Council Directive of 21 May 1991 concerning urban waste water treatment (91/271/EEC). Off. J. Eur. Communities 1991, L135, 40–52. [Google Scholar]
  12. Papp, L.A.; Cardinali-Rezende, J.; de Souza Júdice, W.A.; Sanches, M.B.; Araújo, W.L. Total phosphorus contents currently found in the raw wastewater—Problems and technical solutions for its removal in full-scale wastewater treatment plants. Resour. Conserv. Recycl. 2023, 196, 107026. [Google Scholar] [CrossRef]
  13. Elser, J.; Baker, J.; Boyer, T.; Grieger, K.; Liu, T.; Muenich, R.; Rittmann, B.; Saha, A. Creating an alternative future for Earth’s phosphorus cycle in the Anthropocene via eco-prospecting, eco-mining, and eco-refining. In Earth Systems and Environmental Sciences; Elsevier: Amsterdam, The Netherlands, 2023. [Google Scholar] [CrossRef]
  14. Cordell, D.; Rosemarin, A.; Schröder, J.J.; Smit, A.L. Towards Global Phosphorus Security: A Systems Framework for Phosphorus Recovery and Reuse Options. Chemosphere 2011, 84, 747–758. [Google Scholar] [CrossRef] [PubMed]
  15. Nanda, M.; Kansal, A.; Cordell, D. Managing Agricultural Vulnerability to Phosphorus Scarcity through Bottom-up Assessment of Regional-Scale Opportunities. Agric. Syst. 2020, 184, 102910. [Google Scholar] [CrossRef]
  16. Guaya, D.; Cobos, H.; Camacho, J.; López, C.M.; Valderrama, C.; Cortina, J.L. LTA and FAU-X Iron-Enriched Zeolites: Use for Phosphate Removal from Aqueous Medium. Materials 2022, 15, 5418. [Google Scholar] [CrossRef] [PubMed]
  17. Jupp, A.; Beijer, S.; Narain, G.; Schipper, W.; Slootweg, J. Phosphorus recovery and recycling-closing the loop. Chem. Soc. Rev. 2021, 50, 87–101. [Google Scholar] [CrossRef] [PubMed]
  18. Bergfeldt, B.; Tomasi Morgano, M.; Leibold, H.; Richter, F.; Stapf, D. Recovery of Phosphorus and other Nutrients during Pyrolysis of Chicken Manure. Agriculture 2018, 8, 187. [Google Scholar] [CrossRef]
  19. Quist-Jensen, C.A.; Wybrandt, L.; Løkkegaard, H.; Antonsen, S.B.; Christensen, M.L. Pilot-scale study for phosphorus recovery by sludge acidification and dewatering. Environ. Technol. 2020, 41, 22. [Google Scholar] [CrossRef] [PubMed]
  20. Hollas, C.E.; Bolsan, A.C.; Venturin, B.; Bonassa, G.; Tápparo, D.C.; Cândido, D.; Antes, F.G.; Vanotti, M.B.; Szögi, A.A.; Kunz, A. Second-Generation Phosphorus: Recovery from Wastes towards the Sustainability of Production Chains. Sustainability 2021, 13, 5919. [Google Scholar] [CrossRef]
  21. Krishnamoorthy, N.; Dey, B.; Unpaprom, Y.; Ramaraj, R.; Maniam, G.P.; Govindan, N.; Jayaraman, S.; Arunachalam, T.; Paramasivan, B. Engineering principles and process designs for phosphorus recovery as struvite: A comprehensive review. J. Environ. Chem. Eng. 2021, 9, 105579. [Google Scholar] [CrossRef]
  22. Tang, S.; Liang, J.; Xu, X.; Jin, Y.; Xuan, W.; Li, O.; Fang, L.; Li, Z. Targeting phosphorus transformation to hydroxyapatite through sewage sludge pyrolysis boosted by quicklime toward phosphorus fertilizer alternative with toxic metals compromised. Renew. Sustain. Energy Rev. 2023, 183, 113474. [Google Scholar] [CrossRef]
  23. Pinelli, D.; Bovina, S.; Rubertelli, G.; Martinelli, A.; Guida, S.; Soares, A.; Frascari, D. Regeneration and modelling of a phosphorous removal and recovery hybrid ion exchange resin after long term operation with municipal wastewater. Chemosphere 2022, 286, 1331581. [Google Scholar] [CrossRef] [PubMed]
  24. ESPP, German Phosphorus Platform, Netherlands Nutrient Platform. Catalogue of Phosphorus Recovery Technologies Eur. Sustain. Phosphorus Platf. 2023. Available online: https://phosphorusplatform.eu/images/download/ESPP-NNP-DPP_nutrient-recovery_tech_catalogue.pdf (accessed on 31 March 2024).
  25. Zhu, F.; Cakmak, E.K.; Cetecioglu, Z. Phosphorus recovery for circular Economy: Application potential of feasible resources and engineering processes in Europe. J. Chem. Eng. 2023, 454, 140153. [Google Scholar] [CrossRef]
  26. Egle, L.; Rechberger, H.; Krampe, J.; Zessner, M. Phosphorus recovery from municipal wastewater: An integrated comparative technological, environmental and economic assessment of P recovery technologies. Sci. Total Environ. 2016, 571, 522–542. [Google Scholar] [CrossRef] [PubMed]
  27. Otieno, B.; Funani, C.K.; Khune, S.M.; Kabuba, J.; Osifo, P. Struvite recovery from anaerobically digested waste-activated sludge: A short review. J. Mater. Res. 2023, 38, 3815–3826. [Google Scholar] [CrossRef]
  28. Vali, N.; Combres, A.; Hosseinian, A.; Pettersson, A. The Effect of the Elemental Composition of Municipal Sewage Sludge on the Phosphorus Recycling during Pyrolysis, with a Focus on the Char Chemistry—Modeling and Experiments. Separations 2023, 10, 31. [Google Scholar] [CrossRef]
  29. Ottosen, L.M.; Kirkelund, G.M.; Jensen, P.E.; Pedersen, K.B. Extraction of Phosphorus from Sewage Sludge Ash—Influence of Process Variables on the Electrodialytic Process. Sustainability 2023, 15, 13953. [Google Scholar] [CrossRef]
  30. Fan, X.; Wu, Y.; He, Y.; Liu, H.; Guo, J.; Li, B.; Peng, H. Efficient removal of phosphorus by adsorption. Phosphorus, Sulfur, Silicon Relat. Elem. 2023, 198, 375–384. [Google Scholar] [CrossRef]
  31. Kalaitzidou, K.; Mitrakas, M.; Raptopoulou, C.; Tolkou, A.; Palasantza, P.-A.; Zouboulis, A. Pilot-Scale Phosphate Recovery from Secondary Wastewater Effluents. Environ. Process. 2016, 3, 5–22. [Google Scholar] [CrossRef]
  32. Deng, L.; Dhar, B.R. Phosphorus recovery from wastewater via calcium phosphate precipitation: A critical review of methods, progress, and insights. Chemosphere 2023, 330, 138685. [Google Scholar] [CrossRef] [PubMed]
  33. Fu, J.; Lin, Z.; Zhao, P.; Wang, Y.; He, L.; Zhou, J. Establishment and efficiency analysis of a single-stage denitrifying phosphorus removal system treating secondary effluent. Bioresour. Technol. 2019, 288, 121520. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, C.; Guisasola, A.; Baeza, J.A. A review on the integration of mainstream P-recovery strategies with enhanced biological phosphorus removal. Water Res. 2022, 212, 118102. [Google Scholar] [CrossRef] [PubMed]
  35. Alazaiza, M.Y.D.; He, S.; Su, D.; Abu Amr, S.S.; Toh, P.Y.; Bashir, M.J.K. Sewage Water Treatment Using Chlorella Vulgaris Microalgae for Simultaneous Nutrient Separation and Biomass Production. Separations 2023, 10, 229. [Google Scholar] [CrossRef]
  36. Chrispim, M.C.; Scholz, M.; Nolasco, M.A. Phosphorus recovery from municipal wastewater treatment: Critical review of challenges and opportunities for developing countries. J. Environ. Manag. 2019, 248, 109268. [Google Scholar] [CrossRef] [PubMed]
  37. Carrillo, V.; Fuentes, B.; Gómez, G.; Vidal, G. Characterization and recovery of phosphorus from wastewater by combined technologies. Rev. Environ. Sci. Biotechnol. 2020, 19, 389–418. [Google Scholar] [CrossRef]
  38. Raptopoulou, C.; Kalaitzidou, K.; Tolkou, A.; Palasantza, P.-A.; Mitrakas, M.; Zouboulis, A. Phosphate Removal from Effluent of Secondary Wastewater Treatment: Characterization of Recovered Precipitates and Potential Re-use as Fertilizer. Waste Biomass Valorization 2016, 7, 850–860. [Google Scholar] [CrossRef]
  39. Raptopoulou, C.; Palasantza, P.A.; Mitrakas, M.; Kalaitzidou, K.; Tolkou, A.; Zouboulis, A. Statistical variation of nutrient concentrations and biological removal efficiency of a wastewater treatment plant. Water Util. J. 2016, 14, 5–17. [Google Scholar]
  40. Kalaitzidou, K.; Zouboulis, A.; Mitrakas, M. Thermodynamic Study of Phosphate Adsorption and Removal from Water Using Iron Oxyhydroxides. Water 2022, 14, 1163. [Google Scholar] [CrossRef]
  41. Kalaitzidou, K.; Tolkou, A.; Mitrakas, M.; Zouboulis, A. Phosphorus Removal from Aqueous Solutions and Wastewater samples by Chemical Precipitation. In Proceedings of the IWA Balkan Young Water Professionals, Thessaloniki, Greece, 10–12 May 2015. [Google Scholar]
  42. Palasantza, P.-A.; Germanidis, G.; Tolkou, A.; Mitrakas, M.; Kalaitzidou, K.; Tsimpoukas, N.; Noula, K.; Zouboulis, A. Phosphorus Recovery from Effluent of WWTP in Greece: Design, Built and Operation of a Pilot Plant. In Proceedings of the IWA Specialists Conference on Nutrient Removal and Recovery: Moving Innovation into Practice, Gdansk, Polland, 18–21 May 2015. [Google Scholar]
  43. Tolkou, A.; Raptopoulou, C.; Zouboulis, A.; Kalaitzidou, K.; Mitrakas, M.; Palasantza, P.-A.; Noula, K.; Christodoulou, K. Phosphorus recovery in wastewater treatment: Moving from lab to pilot scale. In Proceedings of the World Congress on New Technologies (NewTech 2015), Barcelona, Spain, 15–17 July 2015. [Google Scholar]
  44. Vu, M.T.; Duong, H.C.; Wang, Q.; Ansari, A.; Cai, Z.; Hoang, N.B.; Nghiem, L.D. Recent technological developments and challenges for phosphorus removal and recovery toward a circular economy. Environ. Technol. Innov. 2023, 30, 103114. [Google Scholar] [CrossRef]
  45. Lu, S.; Pei, L.; Bai, X. Study on method of domestic wastewater treatment through new-type multi-layer artificial wetland. Int. J. Hydrogen Energy 2015, 40, 11207–11214. [Google Scholar] [CrossRef]
  46. Tresintsi, S.; Simeonidis, K.; Pliatsikas, N.; Vourlias, G.; Patsalas, P.; Mitrakas, M. The role of SO42− surface distribution in arsenic removal by iron oxy-hydroxides. J. Solid State Chem. 2014, 213, 145–151. [Google Scholar] [CrossRef]
  47. Amy, G.; Chen, H.W.; Drizo, A.; von Gunten, U.; Brandhuber, P.; Hund, R.; Chowdhury, Z.; Kommineni, S.; Sinha, S.; Jekel, M.; et al. Adsorbent Treatment Technologies for Arsenic Removal; AWWA Research Foundation and American Water Works Association: Washington, DC, USA, 2005. [Google Scholar]
  48. Guo, Y.; Sanjaya, E.H.; Wang, T.; Rong, C.; Luo, Z.; Xue, Y.; Chen, H.; Li, Y.Y. The phosphorus harvest from low-temperature mainstream wastewater through iron phosphate crystallization in a pilot-scale partial nitritation/anammox reactor. Sci. Total Environ. 2023, 862, 160750. [Google Scholar] [CrossRef] [PubMed]
  49. Kalaitzidou, K.; Zouboulis, A.; Mitrakas, M. Cost evaluation for Se(IV) removal, by applying common drinking water treatment processes: Coagulation/precipitation or adsorption. J. Environ. Chem. Eng. 2020, 8, 104209. [Google Scholar] [CrossRef]
  50. Petruzzelli, D.; Dell’Erba, A.; Liberti, L.; Notarnicola, M.; Sengupta, A.K. A phosphate-selective sorbent for the REM NUT® process: Field experience at Massafra Wastewater Treatment Plant. React. Funct. Polym. 2004, 60, 195–202. [Google Scholar] [CrossRef]
  51. Drenkova-Tuhtan, A.; Schneider, M.; Franzreb, M.; Meyer, C.; Gellermann, C.; Sextl, G.; Mandel, K.; Steinmetz, H. Pilot-scale removal and recovery of dissolved phosphate from secondary wastewater effluents with reusable ZnFeZr adsorbent @ Fe3O4/SiO2 particles with magnetic harvesting. Water Res. 2017, 109, 77–87. [Google Scholar] [CrossRef] [PubMed]
  52. Guida, S.; Rubertelli, G.; Jefferson, B.; Soares, A. Demonstration of ion exchange technology for phosphorus removal and recovery from municipal wastewater. Chem. Eng. J. 2021, 420, 129913. [Google Scholar] [CrossRef]
  53. Cheng, M.; Shi, C.; Hao, L.; Wang, X.; Guo, X.; Liu, R.; Hao, X. Sustainable development of phosphorus recovery: From a product perspective. Sustain. Prod. Consum. 2023, 41, 275–290. [Google Scholar] [CrossRef]
  54. Ekman Burgman, L.; Wallsten, B. Should the Sludge Hit the Farm?—How Chemo-Social Relations Affect Policy Efforts to Circulate Phosphorus in Sweden. Sustain. Prod. Consum. 2021, 27, 1488–1497. [Google Scholar] [CrossRef]
  55. Slocombe, S.P.; Zúñiga-Burgos, T.; Chu, L.; Wood, N.J.; Camargo-Valero, M.A.; Baker, A. Fixing the Broken Phosphorus Cycle: Wastewater Remediation by Microalgal Polyphosphates. Front Plant Sci. 2020, 11, 982. [Google Scholar] [CrossRef] [PubMed]
  56. Geissler, B.; Hermann, L.; Mew, M.C.; Steiner, G. Striving Toward a Circular Economy for Phosphorus: The Role of Phosphate Rock Mining. Minerals 2018, 8, 395. [Google Scholar] [CrossRef]
  57. Kalaitzidou, K.; Pagona, E.; Zouboulis, A.; Mitrakas, M. Exploitation of the fine rejected run of mine (ROM 0–4 mm) material to produce refractories in combination with the mining by-products of magnesite mine. Mater. Chem. Phys. 2022, 292, 126743. [Google Scholar] [CrossRef]
  58. Westerhoff, P.; Highfield, D.; Badruzzaman, M.; Yoon, Y. Rapid small-scale column tests for arsenate removal in iron oxide packed bed columns. J. Environ. Eng. 2005, 131, 262–271. [Google Scholar] [CrossRef]
  59. Crittenden, J.C.; Reddy, P.S.; Arora, H.; Trynoski, J.; Hand, D.W.; Perram, D.L.; Summers, R.S. Predicting GAC Performance with Rapid Small-Scale Column Tests. J. AWWA 1991, 83, 77–87. [Google Scholar] [CrossRef]
Figure 1. Comparison of phosphates’ adsorption efficiency using several adsorbents in NSF water matrix at pH 7 and 20 °C.
Figure 1. Comparison of phosphates’ adsorption efficiency using several adsorbents in NSF water matrix at pH 7 and 20 °C.
Water 16 01527 g001
Figure 2. pH effect on phosphate removal for the initial concentration 10 mg PO43−/L and after the addition of (a) 50 mg Ca(OH)2/L, (b) 100 mg Ca.Mg3(CO3)4/L, (c) 75 mg MgO/L, (d) 7.5 mg Fe3+/L (from FeClSO4), and (e) 5 mg Al3+/L (from NaAlO2).
Figure 2. pH effect on phosphate removal for the initial concentration 10 mg PO43−/L and after the addition of (a) 50 mg Ca(OH)2/L, (b) 100 mg Ca.Mg3(CO3)4/L, (c) 75 mg MgO/L, (d) 7.5 mg Fe3+/L (from FeClSO4), and (e) 5 mg Al3+/L (from NaAlO2).
Water 16 01527 g002
Figure 3. Temperature effect on phosphate removal for the initial concentration 5 mg PO43−/L and after the addition of (a) 75 mg Ca(OH)2/L at pH 10, (b) 100 mg CaMg3(CO3)4/L at pH 10, (c) 10 mg MgO/Lat pH 10, (d) 5 mg Fe3+/L (from FeClSO4) at pH 7, and (e) 3 mg Al3+/L (from NaAlO2) at pH 7.
Figure 3. Temperature effect on phosphate removal for the initial concentration 5 mg PO43−/L and after the addition of (a) 75 mg Ca(OH)2/L at pH 10, (b) 100 mg CaMg3(CO3)4/L at pH 10, (c) 10 mg MgO/Lat pH 10, (d) 5 mg Fe3+/L (from FeClSO4) at pH 7, and (e) 3 mg Al3+/L (from NaAlO2) at pH 7.
Water 16 01527 g003
Figure 4. The pilot-plant plc operation; (a) pre-treatment tank with hollow-fiber commercial membranes for the preliminary TSS removal, (b) the adsorption–desorption column, (c) the solution tanks (containing NaOH and CaCl2), and (d) the precipitation tank.
Figure 4. The pilot-plant plc operation; (a) pre-treatment tank with hollow-fiber commercial membranes for the preliminary TSS removal, (b) the adsorption–desorption column, (c) the solution tanks (containing NaOH and CaCl2), and (d) the precipitation tank.
Water 16 01527 g004
Figure 5. Adsorption capacity of phosphorus for the pilot-plant experiments after the subsequent operation cycles.
Figure 5. Adsorption capacity of phosphorus for the pilot-plant experiments after the subsequent operation cycles.
Water 16 01527 g005
Figure 6. Collected Ca-phosphate precipitate: (a) collection in basket filters and (b) dried fertilizer.
Figure 6. Collected Ca-phosphate precipitate: (a) collection in basket filters and (b) dried fertilizer.
Water 16 01527 g006
Table 1. Main physico-chemical characteristic of the iron-based adsorbents used for phosphorus removal. Material from Kalaitzidou, K.; Mitrakas, M.; Raptopoulou, C.; Tolkou, A.; Palasantza, P.-A.; and Zouboulis, A. Pilot-Scale Phosphate Recovery from Secondary Wastewater Effluents. Environ. Process. (Springer Nature) published in 2016 reproduced with permission from SNCSC [31].
Table 1. Main physico-chemical characteristic of the iron-based adsorbents used for phosphorus removal. Material from Kalaitzidou, K.; Mitrakas, M.; Raptopoulou, C.; Tolkou, A.; Palasantza, P.-A.; and Zouboulis, A. Pilot-Scale Phosphate Recovery from Secondary Wastewater Effluents. Environ. Process. (Springer Nature) published in 2016 reproduced with permission from SNCSC [31].
AdsorbentFe (wt.%)Mn (wt.%)BET 1 (m2/g)IEP 2 (mV)ZPC 3 (mV)Ctotal
(mmol [OH]/g)
FeOOH44.9537.22.93.2
AquAsZero3811.51877.353.22.7
GEH54.22377.155.20.9
Bayoxide521357.47.80.3
Note: 1 Brunauer–Emmett–Teller; 2 Iso-Electric Point; and 3 Zero Point Charge.
Table 2. Main chemical characteristics (parameters) of Thessaloniki’s tap (potable) water, NSF water, and of the WWTP AINEIA secondary treated effluent used in the experiments.
Table 2. Main chemical characteristics (parameters) of Thessaloniki’s tap (potable) water, NSF water, and of the WWTP AINEIA secondary treated effluent used in the experiments.
ParameterPotableNSF WaterWWTP Treated Effluent
pH7.5 ± 0.19.2 17.4 ± 0.3
Conductivity μS/cm590 ± 10 6954400 ± 200
Hardness mg CaCO3/L30 ± 115.2750 ± 50
HCO3 mg/L341.6 ± 5138500 ± 50
NO2 mg/LNDND0.45 ± 0.1 m
ΝH4+ mg/LNDND0.6 ± 0.2
PO43− mg/L0.0550.0412–17
Ca2+ mg/L50.2 ± 240120 ± 10
Mg2+ mg/L28.1 ± 112.7100 ± 10
Na+ mg/L8.1 ± 0.588.8585 ± 10
K+ mg/L2.4 ± 0.5ND37 ± 5
Notes: 1 According to the requirements of the respective experiment the pH values of NSF water adjusted respectively.
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

Kalaitzidou, K.; Mitrakas, M.; Zouboulis, A. Post-Removal of Phosphorus from Biologically Treated Wastewater and Recovering It as Fertilizer: Pilot-Scale Attempt—Project PhoReSe. Water 2024, 16, 1527. https://doi.org/10.3390/w16111527

AMA Style

Kalaitzidou K, Mitrakas M, Zouboulis A. Post-Removal of Phosphorus from Biologically Treated Wastewater and Recovering It as Fertilizer: Pilot-Scale Attempt—Project PhoReSe. Water. 2024; 16(11):1527. https://doi.org/10.3390/w16111527

Chicago/Turabian Style

Kalaitzidou, Kyriaki, Manassis Mitrakas, and Anastasios Zouboulis. 2024. "Post-Removal of Phosphorus from Biologically Treated Wastewater and Recovering It as Fertilizer: Pilot-Scale Attempt—Project PhoReSe" Water 16, no. 11: 1527. https://doi.org/10.3390/w16111527

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