Next Article in Journal
The Existence Theorems of Fractional Differential Equation and Fractional Differential Inclusion with Affine Periodic Boundary Value Conditions
Next Article in Special Issue
Photosynthetic Microorganisms: Cultivation and Application
Previous Article in Journal
Reversible Data Hiding in Encrypted Images Based on an Adaptive Recognition Strategy for Blocks
Previous Article in Special Issue
Impact of Light Stress on the Synthesis of Both Antioxidants Polyphenols and Carotenoids, as Fast Photoprotective Response in Chlamydomonas reinhardtii: New Prospective for Biotechnological Potential of This Microalga
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Wastewater Treatment Using Photosynthetic Microorganisms

by
Cristian A. Sepúlveda-Muñoz
1,2,
Ignacio de Godos
1,3 and
Raúl Muñoz
1,2,*
1
Institute of Sustainable Processes, Dr. Mergelina s/n, 47011 Valladolid, Spain
2
Department of Chemical Engineering and Environmental Technology, School of Industrial Engineering, University of Valladolid, Dr. Mergelina s/n, 47011 Valladolid, Spain
3
Department of Chemical Engineering and Environmental Technology, School of Forestry, Agronomic and Bioenergy Industry Engineering (EIFAB), University of Valladolid, Campus Duques de Soria, 42004 Soria, Spain
*
Author to whom correspondence should be addressed.
Symmetry 2023, 15(2), 525; https://doi.org/10.3390/sym15020525
Submission received: 18 January 2023 / Revised: 9 February 2023 / Accepted: 13 February 2023 / Published: 16 February 2023
(This article belongs to the Special Issue Photosynthetic Microorganisms: Cultivation and Application)

Abstract

:
Wastewaters are mainly classified as domestic, industrial and agro-industrial based on their production source. Piggery wastewater (PWW) is a livestock wastewater characterized by its high concentrations of organic matter and ammonium, and by its odour nuisance. Traditionally, PWW has been treated in open anaerobic lagoons, anaerobic digesters and activated sludge systems, which exhibit high greenhouse gas emissions, a limited nutrients removal and a high energy consumption, respectively. Photosynthetic microorganisms can support a sustainable wastewater treatment in engineered photobioreactors at low operating costs and with an efficient recovery of carbon, nitrogen and phosphorous. These microorganisms are capable of absorbing solar irradiation through the photosynthesis process to obtain energy, which is used for their growth and associated carbon and nutrients assimilation. Purple phototrophic bacteria (PPB) represent the photosynthetic microorganisms with the most versatile metabolism in nature, whereas microalgae are the most-studied photosynthetic microorganisms in recent years. This review describes the fundamentals, symmetry and asymmetry of wastewater treatment using photosynthetic microorganisms such as PPB and microalgae. The main photobioreactor configurations along with the potential of PPB and microalgae biomass valorisation strategies are also discussed.

1. Introduction

The human population in the world reached 8 billion in 2022 and recent projections suggested a population up to 9.7 billion by 2050, which represents future challenges for the sustainable development of mankind [1]. This demographic expansion will entail global changes generated by an intensive anthropogenic activity on the Earth. The development of new technologies that reduce pollution and pave the way towards a sustainable development at a global scale represents a severe challenge nowadays. In this context, multiple technologies have been engineered in recent years to reduce the pollution generated by anthropogenic activity with an extensive innovation in wastewater treatment technology [2,3].
Wastewaters are mainly classified depending on their origin as domestic, industrial and agro-industrial [4]. Depending on their origin, wastewaters exhibit a different physical, chemical and microbiological composition [5]. Domestic wastewaters are generated from human activities and characterized by low concentrations of organic matter and nutrients (Table 1) [6]. On the other hand, industrial wastewaters are one of the most diverse sources of wastewaters and their compositions are industry specific. For instance, brewery wastewaters are mainly composed of high concentrations of carbon (sugars and alcohols) [7], whereas wastewater generated by the mining industry contains highly toxic and recalcitrant pollutants such as heavy metals. Finally, agro-industrial wastewaters are produced by intensive agriculture, animal husbandry and their processing. Of them, wastewaters from husbandry of cows, poultry, pigs, turkeys and fishes rank among the most pernicious as a result of their large volumes of production and the gradual increase in animal production. Livestock wastewaters are identified due to the elevated content of carbon from animal manure and a high content of ammonium from animal urine (Table 1) [8,9].

2. Piggery Wastewater

The increased livestock farming for human food production represents a big environmental problem in regions such as Europe, Asia and North America, where the highest densities of animal farms are located. Pigs production in the EU amounted 23,720 thousand metric tons of pig meat in 2021, with Spain as the second largest pork producer in the EU [10]. Other countries such as China and United States produced 47,500 and 12,559 thousand metric tons of pig meat in 2021, respectively (United States, Department of Agriculture). Intensive animal husbandry generates large volumes of wastewaters with a significant content of carbon, nitrogen and phosphorous, which requires the decentralized implementation of cost-effective wastewater treatment. Indeed, the uncontrolled discharge of pig manure entails a high impact in the environment due to the massive greenhouse gas emissions from anaerobic fermentation of pig manure and the released nitrogen and phosphorous, which negatively impacts biodiversity [3,8,11,12]. Piggery wastewater (PWW) is considered one the most polluting agro-industrial wastewaters [8], characterized by high concentrations of organic matter, volatile fatty acids (carbon source), NH4+ (nitrogen source) and even presence of heavy metals (such as copper, lead and zinc) [8,13,14]. In addition, the inadequate management of this wastewater can cause malodour pollution in the surrounding areas.
Depending on the farming practices and type of pig farmed (litters, piglet or sows), PWW concentration can vary due to differences in animal nutrition, farming practices and water usage in cage washing and maintenance (Table 2). Overall, these wastewaters are identified due to elevated concentration of dissolved carbon, nitrogen and phosphorus, and suspended solids.
On the other hand, drugs (such as antiparasitics and antibiotics) are intensively applied during pigs farming to maintain animal health. This represents a serious problem because it is estimated that between 30–90% of the supplemented dugs are excreted without change in their chemical structure, promoting high concentrations of drugs in PWW [15]. Traditionally, these compounds are referred to as emerging contaminants and entail potential risks to the environment and humans [16]. Emerging contaminants such as the antiparasitic fenbendazole at concentrations of 396 ng L−1, the antibiotics of the β-lactams family such as penicillin G (up to 0.03 mg L−1), tetracyclines such as doxycycline (up to 0.16 mg L−1), fluoroquinolones such as enrofloxacin (up to 1.65 mg L−1) and sulfonamides such as sulfadimidine (up to 0.06 mg L−1) have been detected in PWW [17]. These emerging pollutants can be released to the environment if the PWW is not properly treated, thus triggering the risk of pathogenic bacteria evolution on antibiotics resistance.
Table 2. Pollutant concentrations in PWW from different farm around the world.
Table 2. Pollutant concentrations in PWW from different farm around the world.
FarmCOD (mg L−1)TN (mg L−1)TP (mg L−1)TSS (mg L−1)Reference
Daejeon
(South Korea)
18,700810290-[18]
Barcelona
(Spain)
74507851203100[19]
Yokohama
(Japan)
53001270--[20]
Castilla y León
(Spain)
54,00050001500-[8]
Seosan-si
(South Korea)
8420115034-[21]
Santiago
(Chile)
18,4001085172-[22]
Queensland
(Australia)
413011601602420[3]

3. Conventional Piggery Wastewater Treatment Technologies

Multiple technologies have been proposed and investigated for the treatment and valorisation of PWW. Traditionally, PWW has been treated mainly by aerobic or anaerobic processes, and even by a combination of thereof. One the most commonly implemented approaches for PWW management is the use of single-stage or multi-stage anaerobic lagoons to store and treat PWW [23]. This approach is responsible for the release to the open atmosphere of high concentrations of greenhouse gases such as CO2, CH4, NH3 and H2S, with several environmental adverse effects. The decanted manure from these anaerobic lagoons is typically transformed into a biofertilizer with a high concentration of nitrogen and phosphorous, despite the fact it might contain antibiotics and heavy metals that accumulate and build-up in the food chain. In addition, these effluents contain a significant concentration of pathogens that are released in areas where manure is spread, thus polluting surface water and groundwater [24].
Another conventional technology for PWW treatment is anaerobic digestion (AD). This bioprocess is carried out by bacteria and archaea in the absence of O2 or oxidized forms of nitrogen. Those microorganisms degrade the organic matter present in PWW producing mainly methane (CH4) at concentrations ranging from 40 to 75%, carbon dioxide (CO2) at concentrations of 15 to 60% and other gases at lower concentrations such as hydrogen (H2 > 1%), ammonium (NH3 > 1%), oxygen (O2 > 1%) and hydrogen sulphide (H2S > 3%) [25,26,27]. This type of technology allows the conversion of PWW into a renewable gas energy product, namely biogas. However, AD is capable of removing the carbon present in PWW, generating liquid effluents with high concentrations of dissolved nitrogen and phosphorus, which are not assimilated by the anaerobic microbial community during this process. Additionally, this process generates solid sludge corresponding mainly to the non-degradable part of PWW and the anaerobic biomass formed in the process, which must be disposed of in landfills prior to composting.
Activated sludge treatment is an aerobic process that is also typically used for PWW treatment despite its high energy demand [28]. This process consists of an initial removal of suspended solids followed by an aerobic bioreactor with activated sludge, and a final step of settling to collect the biomass generated [29]. The aerobic tank, where carbon is oxidized to CO2 and H2O, and ammonium to nitrite or nitrate, can be preceded by an anoxic tank devoted to heterotrophic denitrification (NO3 to N2). Activated sludge systems are dominated by bacteria, although eukaryotes (protozoa and fungi), archaea and viruses are typically found in this type of process. This process is the most studied and extensively implemented worldwide for the treatment of domestic wastewaters. Unfortunately, this process needs air supplementation via mechanical aeration for an adequate oxygenation by means of blowers or turbines, which involves high operating costs due to the associated high energy demand. In addition, another limitation of activated sludge processes derives from the fact that carbon and nitrogen removal is carried out via volatilization into the atmosphere. Indeed, the organic matter present in PWW is converted into CO2, which contributes to the production of greenhouse gases, whereas nitrogen is converted into N2 through nitrification and denitrification processes or volatilize as NH3. Finally, the heavy metals such as copper and zinc present in PWW are retained in the biomass generated, which is discarded in landfills.

4. Photosynthetic Piggery Wastewater Treatment

The high environmental impacts, and high operating and energy costs of conventional anaerobic and aerobic treatments used for PWW treatment have promoted research on photosynthetic PWW treatment in recent years [3,8,12,30,31,32,33]. Photosynthetic microorganisms are mainly classified in two groups according to their unique metabolism. The first group is composed of purple phototrophic bacteria (PPB), microorganisms that perform anoxygenic photosynthesis (without oxygen production) (Figure 1). This group is subdivided into two bacterial group: purple sulphur bacteria (PSB) and purple non-sulphur bacteria (PNSB) [34]. PPB were originally classified based on their sulphide tolerance and characteristic purple–red colour, although PPB were recently classified according to its class within the phylum Proteobacteria, where PSB belong to the class Gammaproteobacteria and PNSB belong to the class Alphaproteobacteria or Betaproteobacteria [35]. Another photosynthetic bacterial group capable of performing anoxygenic photosynthesis are green sulphur bacteria (GSB) and green non-sulphur bacteria (GNSB), classified in these groups according to their sulphur tolerance and their characteristic green colour [27].
On the other hand, microalgae are a diverse group of photosynthetic microorganisms with eukaryotic and prokaryotic species living in marine, freshwater and terrestrial habitats capable of performing oxygenic photosynthesis (oxygen production). The major classification of eukaryotic microalgae is subdivided in Chlorophyta, Rhodophyta and Phaeophyta, which were originally described according to their characteristic colour (green, red and brown, respectively) [36]. Another group of eukaryotic microalgae are diatoms, which are characterized by cell walls composed of two siliceous pieces and classified into three classes: Coscinodiscophyceae, Mediophyceae and Bacillariophyceae [37]. In addition, prokaryotic microalgae belonging to the phylum Cyanobacteria have evolved in a wide variety of freshwater and saltwater environments.
PPB and microalgae possess extraordinary metabolic capabilities, and are able to grow at low temperatures or extremely high temperatures, in different pH ranges (alkaline or acidic), at high salt concentrations and even under the presence or absence of oxygen [9,35,38,39].

4.1. PPB-Based Treatment

PPB are gram-negative bacteria, prokaryotes and have the most versatile metabolism among microorganisms (Figure 2). Rhodopseudomonas palustris, Rhodobacter sphaeroides, Rhodospirillum rubrum from the PNSB group, and Allochromatium vinosum and Thiocapsa roseopersicina corresponding from the PSB group ranked among the most investigated PPB [35,40]. PNSB are capable of growing under anaerobic and aerobic conditions via phototrophic and chemotrophic metabolisms, respectively [14,41,42]. In addition, PPB are capable of assimilating and growing using organic or inorganic compounds as a carbon source under heterotrophic and autotrophic metabolism, respectively. Some PPB species can obtain energy for growth from the oxidation of inorganic compounds chemolithotrophically using H2 or S2O3−2 as electron donors [35]. PPB can grow in environments with extreme conditions: high concentrations of salt [43], extremely high temperatures or low temperatures in the range of 10–13 °C [9,44,45] and very acid or alkaline conditions [38].
PPB can obtain energy though of anoxygenic photosynthesis process, absorbing energy from the sunlight radiation reaching the earth surface by photon absorption using photosynthetic complexes. In this process, PPB produce ATP via photophosphorylation under phototrophic growth [34,35]. Indeed, PPB are more efficient photosynthetic microorganisms in the photoconversion compared to microalgae [6]. The first descriptions of the basic mechanisms underlying photosynthesis were based on investigation using PPB as model microorganisms [47]. Electron flow during the photosynthetic process in PPB begins with photons absorption by multiple light-harvesting (LH) complexes composed of pigments (mainly of bacteriochlorophylls and carotenoids) bonded to proteins. The absorbed energy is conducted to the reaction centre (P870), where photosynthetic electron transport reactions begin. When P870 is excited (representation P870*, Figure 3), it becomes a very strong electron donor, capable of transferring energize electrons to a quinone (Q) located within the cytoplasmic membrane. These electrons, in turn, are donated to the cytochrome bc1, which translocates four protons for every two electrons from the reaction centre. This process entails an increase in the proton concentration in the periplasm of PPB, which triggers the proton motive force to the cytoplasm, where ATPase enzyme synthetizes adenosine triphosphate (ATP) from adenosine diphosphate (ADP) and inorganic phosphate. In contrast to oxygenic photosynthesis (a process performed by microalgae), the electrons from cytochrome in PPB are not transferred to an electron acceptor (such as molecular oxygen) but to cytochrome C2, and return to the reaction centre P870, closing the cyclic photophosphorylation in PPB [39].
PPB are a unique group of microorganisms capable of absorbing radiation of wavelengths in the near-infrared (NIR) spectrum (Figure 4). Indeed, PPB exhibit a special spectral niche among photosynthetic microorganisms due to the presence of bacteriochlorophyll a and b, which present a maximum absorption at 804 nm and 1040 nm, respectively [48]. PPB contain carotenoids that absorb radiation below 500 nm and provide its characteristic purple–red colour. In addition, carotenoids are able to reduce the photodegradation of bacteriochlorophyll under high irradiance conditions [9].
On the other hand, PPB can grow aerobically under chemoheterotrophic or chemoautotrophic metabolisms, obtaining energy via oxidative phosphorylation (Figure 2) by degrading macromolecules for ATP synthesis [49,50]. Molecular oxygen is used an electron acceptor under these growth conditions. However, this O2 causes a repression and inhibition of bacteriochlorophylls synthesis, which harms the photosynthetic capacity of PPB [51].
R. palustris is a PPB that holds multiple metabolic pathways like the tricarboxylic acid cycle, Embden–Meyerhof pathway, pentose phosphate route, fatty acids metabolism and even can fix CO2 via the Calvin–Benson–Bassham (CBB) cycle [41], which confers this microorganism an efficient metabolism for the assimilation of alcohols, amino acids, carbohydrates, fatty acids, organic acid and even recalcitrant pollutants in wastewater [9,34]. It has been recently described that R. palustris can re-assimilate the carbon produced in the catabolic pathways and strive for maximum carbon efficiency as most CO2 produced from acetate oxidation is fixed and used for biomass production [52]. In addition, PPB can assimilate all forms inorganic of nitrogen such as NO3, NO2, N2 and NH4+ [53]. This ability to fix N2 is remarkable due to the high energy contained in the triple bond of this molecule. In this context, PPBs are able to synthesize the enzyme nitrogenase, which breaks the triple bond present in the N2 and generates NH3 using the energy contained in ATP. PPB can also assimilate the organic nitrogen present in amino acids such as aspartate, glutamate and glutamine. Certain species of PPB can even assimilate aromatic compounds such as cyclohexane carboxylate, toluene, benzoate and their derivatives [35]. On the other hand, due to their extraordinary metabolism PPB are able to remove emerging contaminants such as penicillin G up to 68%, doxycycline up to 38% and enrofloxacin removal up to 88% [17]. These versatile metabolic pathways provide PPB a great potential for the effective treatment of different types of wastewaters.
PPB can treat a wide variety of wastewaters from domestic, industrial and agro-industrial sources with high pollutants removal efficiencies [34,53]. For instance, concomitant removal efficiencies of 63% for COD, 99% for NH4-N and 88% PO4-P have been recorded during domestic wastewater treatment using PPB in batch tests with acetate supplementation [6]. In addition, domestic wastewater treatment by PPB under low radiation intensities (<3 W m−2) with high removal efficiencies over 90% for COD and up to 86% and 91% for TN and TP removal, respectively, has been reported in the literature [54]. In addition, an effective domestic wastewater was achieved in a novel continuous photoanaerobic membrane bioreactor with a PPB dominance in the culture broth over 60% [45]. Similarly, high COD removal efficiencies have been reported during industrial wastewater treatments with PPBs: COD removal of 96% in brewery wastewater [55], 90% in acidic food wastewater [56] and 89% in VFA-rich food industry wastewater [57]. On the other hand, PPB can also support satisfactory organic matter and nutrient removals during agro-industrial wastewaters treatment. For instance, PPB provided removal efficiencies higher than 90% for COD and TN during poultry processing wastewater treatment [32] and higher than 83% for COD during digested piggery wastewater treatment [58]. In addition, PPB exhibit a remarkable metabolism to remove carbon from PWW, with removal efficiencies ranging from 71% to 99% (Table 3). This is mainly due to the fact that the main carbon source present in PWW are volatile fatty acids (VFAs), where CH3COOH, CH3-CH2-CH2-COOH, CH3-CH2-COOH and (CH3)2CHCO2H typically represent ~80% of the dissolved carbon present in PWW [9,59]. PPB can assimilate these VFAs present in PWW via the TCA cycle and finally convert them into cell biomass [60,61]. Although, in open photobioreactors, a fraction of the VFAs present in PWW can be lost via volatilization to the open atmosphere, VFAs assimilation in the form of PPB biomass is the main mechanism of carbon assimilation in closed systems [59]. In addition, PWW treatment can be coupled to biogas upgrading [31]. On the other hand, PPB can remove a limited nitrogen fraction with removal efficiencies of 13% to 42% from PWW via assimilation in the form of PPB biomass (Table 3). However, higher nitrogen removal efficiencies (83–99%) have been recorded in open photobioreactors due to the open nature of these photobioreactors, favouring NH3 volatilization as the main nitrogen removal mechanism. Research on PWW treatment using PPB has been mainly carried out on a small scale, with pioneering research by Hülsen et al. [3] using a flat plate photobioreactor of 60, 80 and 100 L. In addition, enclosed raceways of 450 m2 have been constructed within the H2020-BBI-JU project DEEP PURPLE for domestic wastewater treatment, which represents the largest PPB photobioreactor in the World.

4.2. Microalgae-Based Treatment

Microalgae are photosynthetic microorganisms similar to plants by the fact that they carry out the oxygenic photosynthesis process and both present similar pigments belonging to chlorophyll and carotenoids group. Microalgae are the main photosynthetic microorganisms studied for wastewater treatment since the pioneer investigations of Oswald (1988) [62]. In addition, in the past decades a huge research effort was devoted to microalgae-based technologies for biofuel production, wastewater treatment and production of high value-added bioproducts from microalgae biomass [63,64,65,66]. The main microalgae species investigated were Chlorella vulgaris, Chlamydomonas reinhardtii, Phaeodactylum tricornutum and Arthrospira platensis among others. Microalgae can grow photoautotrophically, fixing CO2 or photoheterotrophically by aerobically metabolizing organic compounds. Both metabolisms utilize energy from sunlight as the main energy source. Microalgae use H2O as electron donor during photoautotrophic growth to produce O2 as a by-product of the reaction [39]. Indeed, this photosynthetic oxygen production is crucial to sustain life in the planet because microalgae produce massive amounts of oxygen in the vast oceans and surface freshwater bodies. This ability to produce oxygen has been successfully used for wastewater treatment, where microalgae grow in symbiosis with bacteria. These bacteria use oxygen as an electron acceptor for organic matter and ammonium oxidation and are able to assimilate the ammonium and phosphorous present in wastewater. In addition, microalgae benefit from essential vitamins and CO2 released by pollutant degrading bacteria (Figure 5) [8,67].
Photosynthesis in microalgae in comparison to PPB differs mainly on the fact that microalgae have two different photosystems for harvesting photons from solar irradiation, photosystem I (PSI or P700) and the photosystem II (PSII or P680) [68]. P680 catalysed the main step in oxygenic photosynthesis, where H2O is split into O2 and electrons (Figure 3). This step starts when light energy is absorbed by P680 which donates an electron to pheophytin a (Ph), turning P680 strongly electropositive and with the capacity that accept electrons from H2O. The oxidation of H2O is produced in a water-oxidizing complex catalysed by a Mn4Ca cluster. Electrons are transported through different quinones (QA and QBe) within P680. Those electrons are used to reduce plastoquinone (PQ) to PQH2. Electrons from PQH2 are transferred to cytochrome b6f. From cytochrome b6f, electrons are transferred to P700 through a copper-containing plastocyanin (PC). The absorption of photons by P700 allows to accept electrons from PC. These different steps generate a proton motive force that is used by ATP synthase to produce ATP [68]. This photosynthesis process is noncyclic because the electrons do not return to P680 and are used for the generation of NADPH from NADP+. Nevertheless, electrons can return from PSI to PSII in some species of cyanobacteria via an electron transport chain linked (dashed line in Figure 3) [39].
The main pigments present in the photosystems of microalgae are chlorophylls, which provide them with their characteristic green colour. Chlorophyll a exhibits maximum absorption at 430 nm and 680 nm, whereas chlorophyll b absorbs mainly at 450 nm and 640 nm [39,48]. In addition, microalgae present a variety of carotenoids such as astaxanthin, β-carotene and lutein that absorb mainly at wavelengths between 400 nm and 520 nm (Figure 4). Carotenoids mainly act as accessory pigments in photon absorption. However, they also have a photoprotective function, preventing damage to the photosystems as a result of microalgae exposure to high irradiations [35,69]. On the other hand, cyanobacteria also contain accessory pigments belonging to the phycobiliprotein group, mainly phycocyanin, phycoerythrin and allophycocyanin. Phycocyanin and allophycocyanin are a blue pigments that absorb mainly at 620 nm and 650 nm [5]. Phycoerythrin is a red pigment that absorbs mainly in the range of 550 nm. Overall, most pigments present in microalgae and cyanobacteria absorb radiation in the visible spectrum (Figure 4).
Microalgae have a powerful photoautotrophic metabolism that allow them to grow using solar energy and assimilating inorganic compounds. Microalgae can efficiently fix CO2 via the Calvin cycle as the main carbon source [70], although are also capable of using NaHCO3 or Na2CO3 under photoautotrophic mode. The preferred forms of nitrogen used by microalgae under photoautotrophic mode are ammonium (NH4+) and nitrate (NO3-) [71], which are used for the synthesis of amino acids and proteins. Some cyanobacteria species are able to fix molecular nitrogen (N2) from the atmosphere as the main nitrogen source. Microalgae can grow photoheterotrophically and utilize organic compounds, such as carboxylic acid, glucose and glycerol, as a carbon source, and amino acids, peptone and urea, as a nitrogen source [72]. However, PPB are more efficient than microalgae under photoheterotrophic growth and biomass production photoheterotrophically is more expensive when synthetic organic compounds are used.
Apart from nitrogen assimilation, the symbiotic association between microalgae and bacteria (Figure 5) can support a sequential nitrification and denitrification, which can further boost nitrogen removal during domestic and agro-industrial wastewater treatment. The nitrification process is performed by ammonia oxidizing and nitrite oxidizing bacteria which can transform ammonium (NH4+) into nitrite (NO2) and nitrite into nitrate (NO3), respectively [73]. Heterotrophic denitrification is an anoxic process performed by heterotrophic bacteria that converts NO3 to NO2 and finally to nitrogen gas (N2).
Microalgae in association with bacteria constitute the mainly studied photosynthetic microorganisms for wastewater treatment. In this context, an efficient treatment of domestic wastewater with TOC removals of 88% and TN of 82% was recorded in anoxic-aerobic photobioreactors [74], whereas the treatment of textile wastewater using this innovative configuration achieved TOC removals of 48%, TN removals of 87% and TP removals 57%, along with an efficient decolourization of this wastewater [75]. An efficient treatment of synthetic food waste digestate, with removal efficiencies up to 96% and 84% for TOC and TN, respectively, was recently reported [76]. In addition, multiple studies devoted to agro-industrial wastewater treatment reporting an efficient carbon and nitrogen removal have been conducted with microalgae [8,12,30,31,33]. Typically, carbon removal efficiencies between 41% and 94%, and nitrogen removals ranging from 56% to 93% have been reported in the literature at laboratory, pilot and demo-scale (Table 4). One of the main problems associated to PWW treatment using microalgae is the inhibition by pH and high concentrations of NH4+, which is the main source of nitrogen present in this type of livestock wastewater. This process limitation has been recently addressed using both PPB and microalgae in sequential photobioreactors, where a significant fraction of carbon and nitrogen are initially removed by PPB. The effluent from PPB photobioreactor is treated in a microalgae photobioreactor, where microalgae favour the assimilation of the remaining nitrogen and the removal of total suspended solids [33].
A recent study compared the removal efficiency of emerging contaminants during PWW treatment by microalgae and PPB targeting 19 veterinary drugs. Microalgae exhibited a more efficient biorremediation of oxytetracycline, doxytetracycline, enrofloxacin, sulfadizine, tiamulin, fenbendazole, penicillin G and sufadimidine compared to PPB. In this study, the decrease in the hydraulic retention time in the photobioreactors produced a detrimental effect in emerging contaminants degradation regardless of the photosynthetic microorganisms used [17].

5. Photobioreactors for Wastewater Treatment

The photobioreactors used for the cultivation of photosynthetic microorganisms are typically classified according to their contact with the atmosphere in open or closed photobioreactors. Photobioreactor design criteria involves a high irradiated surface area, effective mixing and good scalability [65]. Multiple shapes and configurations have been described and investigated for the design of these photobioreactors. In this context, the most common configurations for PPB and microalgae culture are open ponds, open raceways, closed tubular and photo-anaerobic membrane bioreactor (Figure 6) [46]. Open configurations exhibit low investment and operational costs compared to closed photobioreactors. However, they are susceptible to contamination by unwanted microorganisms, the conditions inside the photobioreactor are difficult to control and they have high-water losses by evaporation due to their open configuration [14]. On the other hand, the main disadvantage of closed photobioreactors is their high investment costs. Nevertheless, the limited contact with the open atmosphere avoids microbial contamination of the culture broth, enhances the control of physical–chemical parameters and minimizes water losses.
However, both open and closed photobioreactors must maintain appropriate physical–chemical conditions to support an efficient growth of photosynthetic microorganisms. The key environmental parameter for the cultivation of these microorganisms is radiation, which is necessary for the photosynthetic process but requires extensive surface areas for an effective collection. On the other hand, an optimal control of temperature, pH and dissolved oxygen inside photobioreactors favours PPB and microalgae growth.
Light intensity is the main parameter for the cultivation of photosynthetic microorganisms via phototrophy. The electromagnetic spectrum of solar light provides the energy source for the photosynthetic process. In this regard, photosynthetically active radiation (PAR) is the wavelength range suitable to support the growth of photosynthetic organisms (mainly plants and microalgae), corresponding to wavelengths between 400 nm and 780 nm (visible spectrum), where light absorption by chlorophylls and carotenoids occurs (Figure 4). On the other hand, PPB can absorb radiation in both visible and near-infrared (NIR) spectrum (Figure 7). At high radiation intensities and low culture densities, inhibition of photosynthetic activity can occur, resulting in an increase in intracellular carotenoid concentrations or inhibiting growth totally. However, process operation with dense microalgae cultures can be sustained under light intensities of up to 1300 µmol m−2 s−1, which represents the average solar intensity on a clear day [30,31].
Temperature is also a key parameter for cultivation of photosynthetic microorganisms. Psychrophiles are microorganisms that grow in the temperature range between 0 and 15 °C, mesophiles in temperature range between 15 and 45 °C and thermophiles are microorganisms whose growth optimum temperature exceeds 45 °C [39]. PPB and microalgae species corresponding to these three groups have been consistently reported in the literature. For instance, PPB species such as Rhodoferax antarcticus have been found in the Antarctic [80], whereas the microalga Chlamydomonas nivalis is known to be responsible for the red colour in the snow surface [39]. In addition, an effective treatment of domestic wastewater at low temperatures (10–13 °C) using PPB has been observed [9,44,45]. Most common species of PPB and microalgae, such as Rhodopseudomonas palustris or Chlorella vulgaris, grow at moderated temperatures. Finally, PPB such as Thermochromatium tepidum and the cyanobacterium Oscillatoria terebriformis have been found in hot springs growing at temperatures above 55 °C [38]. Most studies in the literature were carried out using photosynthetic microorganisms at ambient temperatures between 20 to 30 °C.
Microorganisms can be classified in terms of their oxygen tolerance in three groups: (i) aerobic microorganisms that can grow and tolerate high oxygen concentrations, (ii) microaerophiles that can grow in the presence of low concentrations of oxygen and (iii) anaerobes that cannot grow in the presence of oxygen. The latter can be classified as strict or obligate anaerobes, when oxygen is a growth inhibitor that can even kill them, and oxygen tolerant, when oxygen has no biological function in their metabolism [39]. For instance, microaerophilic and anaerobic PPB can tolerate dissolved oxygen concentrations up to 1 mg L−1, which are mainly used in the electron transport chain as an electron acceptor in the oxidative phosphorylation process [50]. However, oxygen can be harmful to PPB during wastewater treatment processes because the growth of aerobic chemotrophic bacteria is favoured [81,82]. On the other hand, although microalgae can grow in the presence of oxygen, very high dissolved oxygen concentrations (>25 mg L−1) can cause damage in microalgae cells and inhibit their growth [65]. Fortunately, the oxygen produced by microalgae is rapidly utilized (as electron acceptor) by bacteria for the oxidation of organic matter and NH4 during wastewater treatment (Figure 5) [33]. In this sense, respirometic methods can be used in wastewater treatment to assess photosynthetic activity in microalgae by calculating the variation of dissolved oxygen concentration over time at a constant temperature. These respirometric methods can be used to analyse toxicity or inhibitory effects induced by wastewaters, determine their biodegradability and energy optimisation of the system to supplement oxygen when necessary [83,84,85].

5.1. Open Photobioreactors

Open ponds were the oldest configurations proposed for photosynthetic microorganisms cultivation at industrial scale for microalgae biomass production [86]. This configuration was proposed to mimic the most common growth of microalgae in nature, using natural or artificial lakes for microalgae culture with different depths. These ponds are less expensive to operate and build compared to other photobioreactor configurations at large-scale. However, their open configuration render them more sensitive to the variations in the weather conditions, to microbial contamination and exhibit high water evaporation levels [87]. This first configuration was limited by the small photic zone when microalgae concentration is high and the high costs associated to nutrient homogenization.
The design of open ponds was improved in order to optimize microalgal activity by reducing their depth to favour photosynthetic activity and reducing energy costs for mixing, traditionally referred to as open raceways. This type of photobioreactor consists of a shallow lagoon (deep of 0.2–0.4 m) with multiple channels and mixing via a paddle wheel motorized (Figure 6) [63]. The continuous mixing by a paddle wheel favours the agitation of the culture broth, exposing the cells to constant radiation during the day (under solar irradiation), which favours the photosynthetic process. This agitation mode is simple and requires a low electricity consumption. Multiple studies describing the cultivation of microalgae and even wastewater treatment in open raceways, with moderate-to-high microalgae productivities are available in the literature [63,64,65,66,88,89]. In this context, microalgae biomass concentrations ranging from 55 mg TSS L−1 in winter and up to 625 mg TSS L−1 in summer have been recorded during an study that assessed the seasonal variation in a pilot scale microalgal-bacterial photobioreactor during digestate treatment coupled with biogas upgrading [89] and up to 2640 mg TSS L−1 during PWW treatment in 3 L open photobioreactor [30]. On the other hand, PPB growth in 10-cm-deep open raceways using aerobic PPB cultures (dominated by Rhodobacter capsulatus) in synthetic media has been reported with biomass concentrations of up to 430 mg TSS L−1 [90]. During PWW treatment in open photobioreactors, PPB biomass production nearby 873 mg TSS L−1 has been described [30]. The main disadvantage of open raceways and open photobioreactors are the high evaporation rate and the associated consumption of water to compensate these evaporation losses [46,64].

5.2. Closed Photobioreactors

The design of closed photobioreactors design involves a limited contact of the cultivation broth with the atmosphere. This configuration exhibits a superior performance in terms of control of microorganisms growth, high surface area-to-volume ratio, low water evaporation losses, higher light availability and easier control of operational parameters, which increase photosynthetic biomass productivity compared to their open counterparts (Table 5) [46,86].
The most studied closed photobioreactors for cultivation of photosynthetic microorganisms are horizontal tubular and flat-panel photobioreactors due to their easier scalability. These photobioreactors need colourless material for their construction, such as glass, polymethyl-methacrylate or polyvinyl chloride (PVC), which increases their investment cost [65]. In this sense, flat-panels photobioreactors using bags of colourless high-density polyethylene supported in a metal cage have been proposed as one of the most economical enclosed photobioreactors for the cultivation of microalgae. On the other hand, tubular configurations are easily scalable by increasing the number and length of tubes and exhibit higher light utilization efficiencies compared to flat-panel photobioreactors. Both types of closed configurations favour the availability of solar radiation, enhance photosynthesis and reduce the footprint required for their construction.
In recent years, the most studied photobioreactor configuration for the cultivation of PPB is the photo-anaerobic membrane bioreactor [32,45,93]. This configuration is composed of a rectangular polymethyl-methacrylate photobioreactor with a submerged flat sheet membrane (0.45 µm pore size) which is operated under anaerobic conditions (air-tight system) with internal gas recycling as an agitation strategy [93]. This type of photobioreactor enables an excellent biomass concentration inside of the photobioreactor and provides an effluent with a low concentration of total solids. However, photo-anaerobic membrane bioreactors also present disadvantages such as the formation of biofilms on the internal walls of the photobioreactor and on the membrane, which hinders an adequate suspension of the PPB cultures. Additionally, the membrane module increases both the investment and operating cost of this technology. Otherwise, high biomass concentrations of up to 920 mg TSS L−1 have been recorded during PWW treatment in closed photobioreactor [33].

6. Biomass Valorisation

The potential high productivities of photosynthetic microorganisms (PPB and microalgae) and the inherent presence of high value-added products represent a unique opportunity to turn wastewater treatment based on photosynthetic microorganisms into a profitable process [34,42,94]. PPB biomass contains a high content of pigments (bacteriochlorophylls and carotenoids), single-cell protein, coenzyme Q10, pantothenic acid, amino acid 5-ALA and polyhydroxyalkanoates [53]. On the other hand, microalgae can synthetize high concentrations of carbohydrates, proteins, lipids and pigments (chlorophylls and carotenoids) [26,70]. In addition, the biomass of photosynthetic microorganisms can be used as a feedstock for the production of biofuels such as biodiesel and bioethanol [95], and some species of PPB and cyanobacteria are capable of producing biohydrogen. PPB and microalgae biomass can be also used as a biofertilizer for plants as an N2-fixing rhizosphere and are known to accumulate compounds that are beneficial for plant growth [76,91] (Figure 8).
Pigments: Photosynthetic microorganisms inherently present high concentrations and a broad portfolio of natural pigments as they are essential molecules for obtaining energy from solar radiation [68]. PPB synthesize a variety of pigments mainly of the group of bacteriochlorophylls and carotenoids, which conform to light-harvesting complexes [34]. Bacteriochlorophyll a and b are pigments found mainly PPB; both pigments have a characteristic green colour. Carotenoids provide the characteristic orange to purple colours to the group of PPB [6]. Thus, the major carotenoids synthesized by PPB are lycopene, rhodopsin, spheroidene, spirilloxanthin and their derivatives [35,96]. On the other hand, microalgae are an important source of commercial high-value compounds and can synthetize mostly pigments of the group of chlorophyll and carotenoids [69,97]. This last group of pigments include β-carotene, astaxanthin and lutein [97,98]. In addition, cyanobacteria can synthesize pigments belonging to the group of phycobiliproteins. An important and rare pigment of this last group is phycocyanin, a pigment with a striking blue colour that can absorb at wavelengths of ~620 nm and present promising antibacterial, anticancer, anti-inflammatory and antioxidant activities [5,99]. All those pigments are nowadays commercialized for the production of cosmetics, pharmaceuticals and functional foods.
Single-cell protein: Single-cell proteins, a term coined to refer to proteins from microbial origin, have been proposed as a sustainable substitute of traditional animal or vegetable proteins. Photosynthetic microorganisms can achieve a fast growth compared to plants and animals and could potentially support high protein productivities, even when growing in low nutrient media or wastewater. These PPB and microalgae biomass valorisation strategies are attracting and increasing attention due to the high commercial value of proteins for the production of animal feed or supplements for human nutrition [34]. PPB typically exhibits a protein content between 41% and 86% [100], whereas microalgae protein a content ranges between 28% and 71% [101]. Spirulina platensis is the most cultivated microalgae (cyanobacteria) for commercial purposes due to its high content of protein (up to 70%) and essential amino acids supporting a high nutritional value. Indeed, S. platensis has been authorized for human consumption as a protein supplement [102], and as a feed supplement in livestock and aquaculture [103]. Single-cell protein can be produced independently of the climatic conditions and availability of arable lands, although nowadays their high costs of production limit the widespread commercialization of single-cell protein. In this context, photosynthetic microorganisms can obtain the energy for growth from sunlight, use wastewater as a nutrient and water source and fix nutrient from atmosphere (CO2 or N2), which represents a unique combination to reduce the production costs of single-cell protein [104].
Coenzyme Q10: PPB can synthesize coenzyme Q10 under phototrophic conditions and accumulate similar concentrations to other microorganisms such as Escherichia coli and Rhizobium radiobater [34]. Coenzyme Q10 productions from 3.1 up to 19.1 mg L−1 d−1 by R. sphaeroides and from 13.9 up to 16.4 mg L−1 d−1 by R. sulfidophilus have been previously reported in PPB species of higher coenzyme Q10 production [34]. The main function of this coenzyme in PPB is the transport of electrons in the electron transport chain, playing an essential role to produce ATP during oxidative phosphorylation. The coenzyme Q10 has antioxidant properties and is widely used in the cosmetic and medical industry [105]. Coenzyme Q10 supplementation in the skin can reduce the wrinkle formation. In addition, this compound prevents and causes beneficial effects in patients with diseases such as Parkinson, cardiovascular and hypertension, thus improving cellular bioenergetics and reducing the occurrence of free radicals due to its antioxidant properties [106]. These benefits for human health entail a high economical value of the coenzyme Q10, with current market prices of ~ EUR 500 per kilogram and increasing demand worldwide in the last years [34]. This coenzyme can be efficiently produced in non-toxic wastewater treatment using PPB [107].
Polyhydroxyalkanoates: Polyhydroxyalkanoates are intracellular biopolymers accumulated by multiple photosynthetic microorganisms and bacteria as energy and carbon reserves [108]. These molecules are a type of biopolymer with biodegradable properties and have a high potential for substituting traditional petrochemical plastics such as polypropylene or polyethylene (produced from petroleum derivatives), which are highly harmful to terrestrial and marine biodiversity [109]. The production of polyhydroxyalkanoates in microorganisms using monocultures or mixed cultures requires a limitation of nitrogen or other essential nutrients to promote the accumulation of these biopolymers in PPB and microalgae [34,42]. In this context, intracellular polyhydroxyalkanoates concentrations nearby 225 mg L−1 have been recorded in PPB biomass [42]. Short feast periods or nutrient limitations favour the adaptation of microorganisms to accumulate polyhydroxyalkanoates over time [108]. The production of polyhydroxyalkanoates from wastewaters by photosynthetic microorganisms requires low concentration of nutrients to favour the intracellular accumulation of this biomolecule under excess of carbon supply. The main PPB described in the literature for their high polyhydroxyalkanoates accumulation capacity are Rhodobacter sphaeroides and Rhodobacter capsulatus [34]. Likewise, cyanobacteria such as Synechococcus sp., Nostoc muscorum and Spirulina platensis have also been described as potential candidates to produce polyhydroxyalkanoates [110]. Although more research is still needed to enhance the accumulation and production of polyhydroxyalkanoates in photosynthetic microorganisms, this solar driven technology platform represents an economic alternative to produce bioplastics.
Biofuels: Photosynthetic microorganisms can produce different types of biofuels, either directly synthesized such as biohydrogen or by additional processes utilizing their biomass such as biodiesel, bioethanol and biogas [34]. Biofuels represent a unique opportunity to boost the development of renewable energy technologies and to reduce CO2 emissions from anthropogenic activity [70].
Photosynthetic microorganisms have emerged as potential candidates for biohydrogen production due to their high conversion efficiency and the broad portfolio of substrates used to sustain their growth. Biohydrogen (H2) can be produced directly by PPB (R. palustris, R. sphaeroides and R. rubrum) mainly under a photoheterotrophic growth mode. Biohydrogen is a clean fuel as during its combustion (H2 + O2) only energy and water (H2O) are produced [111]. This gas can be also used in fuel cells providing a very efficient alternative to combustion engines in transportation [112]. The production of H2 in PPB occurs under illumination, anoxygenic conditions and nitrogen limitations due to the synthesis of the enzyme responsible for hydrogen production (nitrogenase) is inhibited in the presence of ammonium [34,113]. Likewise, microalgae and cyanobacteria species such as Chlamydomonas reinhardtii, Chlorella sorokiniana, Synechocystis sp. or Anabaena sp. can produce directly biohydrogen [114].
Biodiesel was proposed as a potential substitute for conventional petroleum-based fuels. Biodiesel can be produced from plant and animal oils or recently from microorganisms, with a special emphasis on the production from microalgae oil. Biodiesel is a mixture of fatty acid alkyl esters produced by a transesterification process [115]. This process includes three sequential steps, where triglycerides are converted in diglycerides, monoglycerides and finally into esters (biodiesel) and glycerol (by-product). The transesterification process involves a triglyceride and a short chain alcohol (e.g., methanol) in the presence of a catalyst (e.g., sodium hydroxide) for synthesis. The triglyceride-to-biodiesel ratio is 1:1 (e.g., 100 g of biodiesel can be obtained from 100 g of triglycerides). The ability of PPB to produce transesterificable oils to produced biodiesel has not been yet investigated. Over the past 15 years, the remarkable ability of microalgae to synthesize high lipid content under nutrient limitation, with some species capable of producing up to 40-75%, has attracted significant interest and research effort for their mass cultivation as a biofuel source [63,87].
Bioethanol is a biofuel produced from the carbohydrates contained in plant biomass or microorganisms [116]. This biofuel is produced through alcoholic fermentation processes of hydrolysed biomass, which produce an ethanol with an identical chemical composition to that of chemical origin, and that also requires a distillation step for its purification [95]. Bioethanol can be used as a biofuel directly or mixed with conventional gasoline. In the past, ethanol production from microalgae biomass has been proposed based on its high carbohydrate content in terms of cellulose and starch, which can be fermented for bioethanol production [117]. In addition, the CO2 produced from ethanol combustion can be recovered as microalgae are capable of fixing CO2, thus neutralizing carbon emissions to the atmosphere and creating a cycling process for carbon utilization [64,118].
Finally, the most studied biofuel produced from photosynthetic microorganisms biomass is methane. Methane can be produced from anaerobic digestion of PPB or microalgae biomass, where the degradation of the organic matter contained in these photosynthetic microorganisms occurs under anaerobic conditions in different steps catalysed by anaerobic bacteria and archaea [4]. The first process in anaerobic digestion is a hydrolysis step, where complex organic matter is converted into simpler molecules such as amino acids, fatty acids and monosaccharides by different microorganisms. These simpler compounds are converted in a second acidogenesis step into volatile fatty acids and alcohols. The latter compounds are biotransformed by acetogenic bacteria during the acetogenic step into acetic acid, hydrogen and carbon dioxide as the main metabolites. Finally, methanogenic archaea transform the acetic acid, carbon dioxide and H2 into biogas during the methanogenic step [119]. Biogas is a mixture of gases mainly composed of high concentrations of methane (40–75%), carbon dioxide (15–60%), hydrogen sulphide (0.005–3%), nitrogen (0–2%) and oxygen (0–1%) [26]. Biogas upgrading to biomethane is necessary to remove unwanted gases such as CO2 and H2S if biogas is used as vehicle fuel or injected into natural gas grids. This biogas upgrading step can be carried out using PPB [27,31] and microalgae [120,121,122] based on the ability of these microorganisms to fix CO2 and support H2S oxidation.
Biofertilizers: Biofertilizers are fertilizers of a biological origin that are capable of promoting growth and the nutritional quality of plants. PPB can be directly used as biofertilizers (active benefit) as these photosynthetic microorganisms can grow in symbiosis with plants in the rhizosphere and are capable of synthesizing and excreting compounds that promote plant growth. For instance, PPB can synthesize the amino acid 5-ALA, which is beneficial to plant growth [53]. 5-ALA can be used as plant growth enhancer and even possesses insecticide and herbicide properties [34,100]. In addition, when PPB are associated with plant roots, they can contribute to heavy metal remediation in soils via accumulation into PPB biomass, which reduces adverse effects of these metals inside plants [123]. In addition, PPB biomass can be used as dried biomass (passive benefit) due to its high nutrient (N and P) content [34,91]. PPB can accumulate phosphate in form of polyphosphates and fix nitrogen from the atmosphere, nutrients that are necessary for plant growth and limited in soils used for intensive agriculture [91]. On the other hand, the merits of microalgae biomass as slow-release biofertilizers have been consistently reported in the literature [124]. Microalgae biomass is also rich in nitrogen, phosphorous and trace elements such as potassium, calcium, magnesium and iron, that can be absorbed by plants when microalgal biomass is supplemented and even increase the nutritional quality of biofertilized fruits [5].

7. Conclusions and Future Prospects

PWW management represents a serious challenge to the sustainability of pig farming due to its high production and pollutant concentrations. Conventional PWW treatments are only partially able to treat this type of wastewater at the expense of high energy consumption, prohibitive greenhouse gas emissions and a waste of resources. Photosynthetic microorganisms such as PPB and microalgae represent a promising solution for PWW valorisation due to their ability to obtain energy from solar radiation, fast growth and high nutrients assimilation. The high versatility of photosynthetic metabolism supports wastewater treatment under aerobic and anaerobic conditions in a wide variety of photobioreactor configurations. Added value bioproducts such as pigments, single-cell protein, biopolymers, biofuels and biofertilizers can be obtained from both PPB and microalgae biomass. However, despite the promising advances achieved in wastewater treatment using photosynthetic microorganisms, the future research needed to move this platform technology to commercial scale should focus on its optimization under outdoor conditions, which will allow evaluating the effects of direct solar irradiation, daily and seasonal changes in light intensities and temperature, changes in PWW composition and microbial contamination on process performance. In addition, the scale-up of continuous photobioreactors treating PWW using photosynthetic microorganisms in relevant environments and the potential of the photosynthetic biomass generated during PWW treatment must be evaluated.

Author Contributions

Conceptualization, C.A.S.-M., I.d.G. and R.M.; resources, C.A.S.-M. and R.M.; writing—original draft preparation, C.A.S.-M.; writing—review and editing, C.A.S.-M., I.d.G. and R.M.; supervision, I.d.G. and R.M.; project administration, C.A.S.-M. and R.M.; funding acquisition, C.A.S.-M. and R.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded from the Regional Government of Castilla y León, the EU-FEDER programme (CLU 2017-09, CL-EI-2021-07 and UIC 071) and CONICYT (PFCHA/DOCTORADO BECAS CHILE/2017–72180211) is gratefully acknowledged.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. United Nations, Department of Economic and Social Affairs, Population Division. World Population Prospects 2019: Highlights; United Nations: New York, NY, USA, 2019; ISBN 978-92-1-148316-1. [Google Scholar]
  2. Cai, T.; Park, S.Y.; Li, Y. Nutrient Recovery from Wastewater Streams by Microalgae: Status and Prospects. Renew. Sustain. Energy Rev. 2013, 19, 360–369. [Google Scholar] [CrossRef]
  3. Hülsen, T.; Stegman, S.; Batstone, D.J.; Capson-Tojo, G. Naturally Illuminated Photobioreactors for Resource Recovery from Piggery and Chicken-Processing Wastewaters Utilising Purple Phototrophic Bacteria. Water Res. 2022, 214, 118194. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, Y.; Cheng, J.J.; Creamer, K.S. Inhibition of Anaerobic Digestion Process: A Review. Bioresour. Technol. 2008, 99, 4044–4064. [Google Scholar] [CrossRef] [PubMed]
  5. Aditya, L.; Mahlia, T.M.I.; Nguyen, L.N.; Vu, H.P.; Nghiem, L.D. Microalgae-Bacteria Consortium for Wastewater Treatment and Biomass Production. Sci. Total Environ. 2022, 838, 155871. [Google Scholar] [CrossRef] [PubMed]
  6. Hülsen, T.; Batstone, D.J.; Keller, J. Phototrophic Bacteria for Nutrient Recovery from Domestic Wastewater. Water Res. 2014, 50, 18–26. [Google Scholar] [CrossRef] [PubMed]
  7. Lu, H.; Zhang, G.; Dai, X.; Schideman, L.; Zhang, Y.; Li, B.; Wang, H. A Novel Wastewater Treatment and Biomass Cultivation System Combining Photosynthetic Bacteria and Membrane Bioreactor Technology. Desalination 2013, 322, 176–181. [Google Scholar] [CrossRef]
  8. Godos, I.d; Blanco, S.; García-Encina, P.A.; Becares, E.; Muñoz, R. Long-Term Operation of High Rate Algal Ponds for the Bioremediation of Piggery Wastewaters at High Loading Rates. Bioresour. Technol. 2009, 100, 4332–4339. [Google Scholar] [CrossRef]
  9. Sepúlveda-Muñoz, C.A.; Torres-Franco, A.F.; de Godos, I.; Muñoz, R. Exploring the Metabolic Capabilities of Purple Phototrophic Bacteria during Piggery Wastewater Treatment. J. Water Process Eng. 2022, 50, 103317. [Google Scholar] [CrossRef]
  10. Jiménez-Ruiz, S.; Laguna, E.; Vicente, J.; García-Bocanegra, I.; Martínez-Guijosa, J.; Cano-Terriza, D.; Risalde, M.A.; Acevedo, P. Characterization and Management of Interaction Risks between Livestock and Wild Ungulates on Outdoor Pig Farms in Spain. Porc. Health Manag. 2022, 8, 2. [Google Scholar] [CrossRef]
  11. Gilbert, M.; Nicolas, G.; Cinardi, G.; Van Boeckel, T.P.; Vanwambeke, S.O.; Wint, G.R.W.; Robinson, T.P. Global Distribution Data for Cattle, Buffaloes, Horses, Sheep, Goats, Pigs, Chickens and Ducks in 2010. Sci. Data 2018, 5, 180227. [Google Scholar] [CrossRef] [Green Version]
  12. Godos, I.d; Vargas, V.A.; Blanco, S.; González, M.C.G.; Soto, R.; García-Encina, P.A.; Becares, E.; Muñoz, R. A Comparative Evaluation of Microalgae for the Degradation of Piggery Wastewater under Photosynthetic Oxygenation. Bioresour. Technol. 2010, 101, 5150–5158. [Google Scholar] [CrossRef]
  13. Chen, G.; Huang, J.; Tian, X.; Chu, Q.; Zhao, Y.; Zhao, H. Effects of Influent Loads on Performance and Microbial Community Dynamics of Aerobic Granular Sludge Treating Piggery Wastewater. J. Chem. Technol. Biotechnol. 2018, 93, 1443–1452. [Google Scholar] [CrossRef]
  14. Sepúlveda-Muñoz, C.A.; Ángeles, R.; de Godos, I.; Muñoz, R. Comparative Evaluation of Continuous Piggery Wastewater Treatment in Open and Closed Purple Phototrophic Bacteria-Based Photobioreactors. J. Water Process Eng. 2020, 38, 101608. [Google Scholar] [CrossRef]
  15. Portela-Monge, C.; Bolado, S.; López-Serna, R.; Jiménez, J.J. Determination of Contaminants of Emerging Concern in Raw Pig Manure as a Whole: Difference with the Analysis of Solid and Liquid Phases Separately. Environ. Sci. Process. Impacts 2022, 87, 2357–2367. [Google Scholar] [CrossRef] [PubMed]
  16. Zhou, J.L.; Yang, L.; Huang, K.X.; Chen, D.Z.; Gao, F. Mechanisms and Application of Microalgae on Removing Emerging Contaminants from Wastewater: A Review. Bioresour. Technol. 2022, 364, 128049. [Google Scholar] [CrossRef] [PubMed]
  17. López-Serna, R.; García, D.; Bolado, S.; Jiménez, J.J.; Lai, F.Y.; Golovko, O.; Gago-Ferrero, P.; Ahrens, L.; Wiberg, K.; Muñoz, R. Photobioreactors Based on Microalgae-Bacteria and Purple Phototrophic Bacteria Consortia: A Promising Technology to Reduce the Load of Veterinary Drugs from Piggery Wastewater. Sci. Total Environ. 2019, 692, 259–266. [Google Scholar] [CrossRef] [PubMed]
  18. Myung, K.K.; Choi, K.M.; Yin, C.R.; Lee, K.Y.; Im, W.T.; Ju, H.L.; Lee, S.T. Odorous Swine Wastewater Treatment by Purple Non-Sulfur Bacteria, Rhodopseudomonas palustris, Isolated from Eutrophicated Ponds. Biotechnol. Lett. 2004, 26, 819–822. [Google Scholar] [CrossRef]
  19. Obaja, D.; MacÉ, S.; Mata-Alvarez, J. Biological Nutrient Removal by a Sequencing Batch Reactor (SBR) Using an Internal Organic Carbon Source in Digested Piggery Wastewater. Bioresour. Technol. 2005, 96, 7–14. [Google Scholar] [CrossRef] [PubMed]
  20. Joo, H.S.; Hirai, M.; Shoda, M. Piggery Wastewater Treatment Using Alcaligenes faecalis Strain No. 4 with Heterotrophic Nitrification and Aerobic Denitrification. Water Res. 2006, 40, 3029–3036. [Google Scholar] [CrossRef]
  21. Lee, S.A.; Lee, N.; Oh, H.M.; Ahn, C.Y. Stepwise Treatment of Undiluted Raw Piggery Wastewater, Using Three Microalgal Species Adapted to High Ammonia. Chemosphere 2021, 263, 127934. [Google Scholar] [CrossRef]
  22. Palominos, N.; Castillo, A.; Guerrero, L.; Borja, R.; Huiliñir, C. Coupling of Anaerobic Digestion and Struvite Precipitation in the Same Reactor: Effect of Zeolite and Bischofite as Mg2+ Source. Front. Environ. Sci. 2021, 9, 581. [Google Scholar] [CrossRef]
  23. Montefiore, L.R.; Nelson, N.G.; Dean, A.; Sharara, M. Reconstructing the Historical Expansion of Industrial Swine Production from Landsat Imagery. Sci. Rep. 2022, 12, 1736. [Google Scholar] [CrossRef]
  24. Campagnolo, E.R.; Johnson, K.R.; Karpati, A.; Rubin, C.S.; Kolpin, D.W.; Meyer, M.T.; Esteban, J.E.; Currier, R.W.; Smith, K.; Thu, K.M.; et al. Antimicrobial Residues in Animal Waste and Water Resources Proximal to Large-Scale Swine and Poultry Feeding Operations. Sci. Total Environ. 2002, 299, 89–95. [Google Scholar] [CrossRef] [PubMed]
  25. Manyi-Loh, C.E.; Mamphweli, S.N.; Meyer, E.L.; Okoh, A.I.; Makaka, G.; Simon, M. Microbial Anaerobic Digestion (Bio-Digesters) as an Approach to the Decontamination of Animal Wastes in Pollution Control and the Generation of Renewable Energy. Int. J. Environ. Res. Public Health 2013, 10, 4390–4417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Ángeles, R.; Arnaiz, E.; Gutiérrez, J.; Sepúlveda-Muñoz, C.A.; Fernández-Ramos, O.; Muñoz, R.; Lebrero, R. Optimization of Photosynthetic Biogas Upgrading in Closed Photobioreactors Combined with Algal Biomass Production. J. Water Process Eng. 2020, 38, 101554. [Google Scholar] [CrossRef]
  27. Struk, M.; Sepúlveda-Muñoz, C.A.; Kushkevych, I.; Muñoz, R. Photoautotrophic Removal of Hydrogen Sulfide from Biogas Using Purple and Green Sulfur Bacteria. J. Hazard. Mater. 2023, 443, 130337. [Google Scholar] [CrossRef]
  28. Montes, N.; Otero, M.; Coimbra, R.N.; Méndez, R.; Martín-Villacorta, J. Removal of Tetracyclines from Swine Manure at Full-Scale Activated Sludge Treatment Plants. Environ. Technol. 2015, 36, 1966–1973. [Google Scholar] [CrossRef]
  29. Suzuki, K.; Waki, M.; Yasuda, T.; Fukumoto, Y.; Kuroda, K.; Sakai, T.; Suzuki, N.; Suzuki, R.; Matsuba, K. Distribution of Phosphorus, Copper and Zinc in Activated Sludge Treatment Process of Swine Wastewater. Bioresour. Technol. 2010, 101, 9399–9404. [Google Scholar] [CrossRef]
  30. García, D.; de Godos, I.; Domínguez, C.; Turiel, S.; Bolado, S.; Muñoz, R. A Systematic Comparison of the Potential of Microalgae-Bacteria and Purple Phototrophic Bacteria Consortia for the Treatment of Piggery Wastewater. Bioresour. Technol. 2019, 276, 18–27. [Google Scholar] [CrossRef] [Green Version]
  31. Marín, D.; Posadas, E.; García, D.; Puyol, D.; Lebrero, R.; Muñoz, R. Assessing the Potential of Purple Phototrophic Bacteria for the Simultaneous Treatment of Piggery Wastewater and Upgrading of Biogas. Bioresour. Technol. 2019, 281, 10–17. [Google Scholar] [CrossRef]
  32. Hülsen, T.; Hsieh, K.; Tait, S.; Barry, E.M.; Puyol, D.; Batstone, D.J. White and Infrared Light Continuous Photobioreactors for Resource Recovery from Poultry Processing Wastewater—A Comparison. Water Res. 2018, 144, 665–676. [Google Scholar] [CrossRef] [PubMed]
  33. Sepúlveda-Muñoz, C.A.; Hontiyuelo, G.; Blanco, S.; Torres-Franco, A.F.; Muñoz, R. Photosynthetic Treatment of Piggery Wastewater in Sequential Purple Phototrophic Bacteria and Microalgae-Bacteria Photobioreactors. J. Water Process Eng. 2022, 47, 102825. [Google Scholar] [CrossRef]
  34. Capson-Tojo, G.; Batstone, D.J.; Grassino, M.; Vlaeminck, S.E.; Puyol, D.; Verstraete, W.; Kleerebezem, R.; Oehmen, A.; Ghimire, A.; Pikaar, I.; et al. Purple Phototrophic Bacteria for Resource Recovery: Challenges and Opportunities. Biotechnol. Adv. 2020, 43, 107567. [Google Scholar] [CrossRef] [PubMed]
  35. Hunter, C.N.; Daldal, F.; Thurnauer, M.C.; Beatty, J.T. The Purple Phototrophic Bacteria; Advances in Photosynthesis and Respiration; Springer: Dordrecht, The Netherlands, 2009; Volume 28, ISBN 978-1-4020-8814-8. [Google Scholar]
  36. Chan, S.S.; Low, S.S.; Chew, K.W.; Ling, T.C.; Rinklebe, J.; Juan, J.C.; Ng, E.P.; Show, P.L. Prospects and Environmental Sustainability of Phyconanotechnology: A Review on Algae-Mediated Metal Nanoparticles Synthesis and Mechanism. Environ. Res. 2022, 212, 113140. [Google Scholar] [CrossRef]
  37. Medlin, L.K. Evolution of the Diatoms: Major Steps in Their Evolution and a Review of the Supporting Molecular and Morphological Evidence. Phycologia 2016, 55, 79–103. [Google Scholar] [CrossRef]
  38. Hallenbeck, P.C. Modern Topics in the Phototrophic Prokaryotes: Environmental and Applied Aspects; Springer International Publishing: Cham, Switzerland, 2017; ISBN 978-3-319-46259-2. [Google Scholar]
  39. Madigan, M.T.; Bender, K.S.; Buckley, D.H.; Sattley, W.M.; Stahl, D.A. Brock Biology of Microorganisms, 16th ed.; Pearson Education: Harlow, UK, 2021; ISBN 978-1-292-40479-0. [Google Scholar]
  40. Puyol, D.; Hülsen, T.; Padrino, B.; Batstone, D.J.; Martinez, F.; Melero, J.A. Exploring the Inhibition Boundaries of Mixed Cultures of Purple Phototrophic Bacteria for Wastewater Treatment in Anaerobic Conditions. Water Res. 2020, 183, 116057. [Google Scholar] [CrossRef]
  41. Larimer, F.W.; Chain, P.; Hauser, L.; Lamerdin, J.; Malfatti, S.; Do, L.; Land, M.L.; Pelletier, D.A.; Beatty, J.T.; Lang, A.S.; et al. Complete Genome Sequence of the Metabolically Versatile Photosynthetic Bacterium Rhodopseudomonas palustris. Nat. Biotechnol. 2004, 22, 55–61. [Google Scholar] [CrossRef] [Green Version]
  42. Touloupakis, E.; Poloniataki, E.G.; Casciana, M.; Ghanotakis, D.F.; Carlozzi, P. Poly-β-Hydroxybutyrate Production by Rhodopseudomonas sp. Grown in Semi-Continuous Mode in a 4 L Photobioreactor. Symmetry 2021, 13, 1609. [Google Scholar] [CrossRef]
  43. Hülsen, T.; Hsieh, K.; Batstone, D.J. Saline Wastewater Treatment with Purple Phototrophic Bacteria. Water Res. 2019, 160, 259–267. [Google Scholar] [CrossRef]
  44. Dalaei, P.; Ho, D.; Nakhla, G.; Santoro, D. Low Temperature Nutrient Removal from Municipal Wastewater by Purple Phototrophic Bacteria (PPB). Bioresour. Technol. 2019, 288, 121566. [Google Scholar] [CrossRef]
  45. Hülsen, T.; Barry, E.M.; Lu, Y.; Puyol, D.; Batstone, D.J. Low Temperature Treatment of Domestic Wastewater by Purple Phototrophic Bacteria: Performance, Activity, and Community. Water Res. 2016, 100, 537–545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Méndez, L.; Sepúlveda-Muñoz, C.A.; del Rosario Rodero, M.; de Godos, I.; Muñoz, R. Decarbonization Potentials Using Photobiological Systems. In Pathways to Water Sector Decarbonization, Carbon Capture and Utilization; IWA Publishing: London, UK, 2022; pp. 143–170. ISBN 9781789061789. [Google Scholar]
  47. Cogdell, R.J.; Isaacs, N.W.; Howard, T.D.; McLuskey, K.; Fraser, N.J.; Prince, S.M. How Photosynthetic Bacteria Harvest Solar Energy. J. Bacteriol. 1999, 181, 3869–3879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Stomp, M.; Huisman, J.; Stal, L.J.; Matthijs, H.C.P. Colorful Niches of Phototrophic Microorganisms Shaped by Vibrations of the Water Molecule. ISME J. 2007, 1, 271–282. [Google Scholar] [CrossRef]
  49. Lu, H.; Zhang, G.; Wan, T.; Lu, Y. Influences of Light and Oxygen Conditions on Photosynthetic Bacteria Macromolecule Degradation: Different Metabolic Pathways. Bioresour. Technol. 2011, 102, 9503–9508. [Google Scholar] [CrossRef]
  50. Lu, H.; Zhang, G.; Dong, S. Quantitative Study of PNSB Energy Metabolism in Degrading Pollutants under Weak Light-Micro Oxygen Condition. Bioresour. Technol. 2011, 102, 4968–4973. [Google Scholar] [CrossRef] [PubMed]
  51. Izu, K.; Nakajima, F.; Yamamoto, K.; Kurisu, F. Aeration Conditions Affecting Growth of Purple Nonsulfur Bacteria in an Organic Wastewater Treatment Process. Syst. Appl. Microbiol. 2001, 24, 294–302. [Google Scholar] [CrossRef] [PubMed]
  52. Navid, A.; Jiao, Y.; Wong, S.E.; Pett-Ridge, J. System-Level Analysis of Metabolic Trade-Offs during Anaerobic Photoheterotrophic Growth in Rhodopseudomonas palustris. BMC Bioinform. 2019, 20, 233. [Google Scholar] [CrossRef] [Green Version]
  53. Lu, H.; Zhang, G.; Zheng, Z.; Meng, F.; Du, T.; He, S. Bio-Conversion of Photosynthetic Bacteria from Non-Toxic Wastewater to Realize Wastewater Treatment and Bioresource Recovery: A Review. Bioresour. Technol. 2019, 278, 383–399. [Google Scholar] [CrossRef]
  54. Dalaei, P.; Bahreini, G.; Nakhla, G.; Santoro, D.; Batstone, D.; Hülsen, T. Municipal Wastewater Treatment by Purple Phototropic Bacteria at Low Infrared Irradiances Using a Photo-Anaerobic Membrane Bioreactor. Water Res. 2020, 173, 115535. [Google Scholar] [CrossRef]
  55. Yang, A.; Zhao, W.; Peng, M.; Zhang, G.; Zhi, R.; Meng, F. A Special Light-Aerobic Condition for Photosynthetic Bacteria-Membrane Bioreactor Technology. Bioresour. Technol. 2018, 268, 820–823. [Google Scholar] [CrossRef]
  56. Liu, S.; Zhang, G.; Zhang, J.; Li, X.; Li, J. Performance, Carotenoids Yield and Microbial Population Dynamics in a Photobioreactor System Treating Acidic Wastewater: Effect of Hydraulic Retention Time (HRT) and Organic Loading Rate (OLR). Bioresour. Technol. 2016, 200, 245–252. [Google Scholar] [CrossRef] [PubMed]
  57. Liu, S.; Daigger, G.T.; Kang, J.; Zhang, G. Effects of Light Intensity and Photoperiod on Pigments Production and Corresponding Key Gene Expression of Rhodopseudomonas palustris in a Photobioreactor System. Bioresour. Technol. 2019, 294, 122172. [Google Scholar] [CrossRef] [PubMed]
  58. Wen, S.; Liu, H.; He, H.; Luo, L.; Li, X.; Zeng, G.; Zhou, Z.; Lou, W.; Yang, C. Treatment of Anaerobically Digested Swine Wastewater by Rhodobacter blasticus and Rhodobacter capsulatus. Bioresour. Technol. 2016, 222, 33–38. [Google Scholar] [CrossRef] [PubMed]
  59. Sepúlveda-Muñoz, C.A.; de Godos, I.; Puyol, D.; Muñoz, R. A Systematic Optimization of Piggery Wastewater Treatment with Purple Phototrophic Bacteria. Chemosphere 2020, 253, 126621. [Google Scholar] [CrossRef] [PubMed]
  60. Alloul, A.; Wuyts, S.; Lebeer, S.; Vlaeminck, S.E. Volatile Fatty Acids Impacting Phototrophic Growth Kinetics of Purple Bacteria: Paving the Way for Protein Production on Fermented Wastewater. Water Res. 2019, 152, 138–147. [Google Scholar] [CrossRef]
  61. Fradinho, J.C.; Oehmen, A.; Reis, M.A.M. Photosynthetic Mixed Culture Polyhydroxyalkanoate (PHA) Production from Individual and Mixed Volatile Fatty Acids (VFAs): Substrate Preferences and Co-Substrate Uptake. J. Biotechnol. 2014, 185, 19–27. [Google Scholar] [CrossRef]
  62. Oswald, W.J. Micro-Algae and Waste-Water Treatment. In Micro-Algal Biotechnology; Borowitzka, M.B.L., Ed.; Cambridge University Press: Cambridge, UK, 1988. [Google Scholar]
  63. Chisti, Y. Biodiesel from Microalgae. Biotechnol. Adv. 2007, 25, 294–306. [Google Scholar] [CrossRef]
  64. Christenson, L.; Sims, R. Production and Harvesting of Microalgae for Wastewater Treatment, Biofuels, and Bioproducts. Biotechnol. Adv. 2011, 29, 686–702. [Google Scholar] [CrossRef]
  65. Muñoz, R.; Guieysse, B. Algal-Bacterial Processes for the Treatment of Hazardous Contaminants: A Review. Water Res. 2006, 40, 2799–2815. [Google Scholar] [CrossRef]
  66. Posadas, E.; Marín, D.; Blanco, S.; Lebrero, R.; Muñoz, R. Simultaneous Biogas Upgrading and Centrate Treatment in an Outdoors Pilot Scale High Rate Algal Pond. Bioresour. Technol. 2017, 232, 133–141. [Google Scholar] [CrossRef] [Green Version]
  67. Nagarajan, D.; Lee, D.-J.; Varjani, S.; Lam, S.S.; Allakhverdiev, S.I.; Chang, J.-S. Microalgae-Based Wastewater Treatment—Microalgae-Bacteria Consortia, Multi-Omics Approaches and Algal Stress Response. Sci. Total Environ. 2022, 845, 157110. [Google Scholar] [CrossRef] [PubMed]
  68. Senge, M.; Ryan, A.; Letchford, K.; MacGowan, S.; Mielke, T. Chlorophylls, Symmetry, Chirality, and Photosynthesis. Symmetry 2014, 6, 781–843. [Google Scholar] [CrossRef] [Green Version]
  69. Faraloni, C.; Lorenzo, T.D.; Bonetti, A. Impact of Light Stress on the Synthesis of Both Antioxidants Polyphenols and Carotenoids, as Fast Photoprotective Response in Chlamydomonas reinhardtii: New Prospective for Biotechnological Potential of This Microalga. Symmetry 2021, 13, 2220. [Google Scholar] [CrossRef]
  70. Mountourakis, F.; Papazi, A.; Kotzabasis, K. The Microalga Chlorella vulgaris as a Natural Bioenergetic System for Effective CO2 Mitigation-New Perspectives against Global Warming. Symmetry. 2021, 13, 997. [Google Scholar] [CrossRef]
  71. Silveira, C.F.; de Assis, L.R.; de Sousa Oliveira, A.P.S.; Calijuri, M.L. Valorization of Swine Wastewater in a Circular Economy Approach: Effects of Hydraulic Retention Time on Microalgae Cultivation. Sci. Total Environ. 2021, 789, 147861. [Google Scholar] [CrossRef]
  72. Perez-Garcia, O.; Escalante, F.M.E.; De-Bashan, L.E.; Bashan, Y. Heterotrophic Cultures of Microalgae: Metabolism and Potential Products. Water Res. 2011, 45, 11–36. [Google Scholar] [CrossRef]
  73. Chai, W.S.; Chew, C.H.; Munawaroh, H.S.H.; Ashokkumar, V.; Cheng, C.K.; Park, Y.K.; Show, P.L. Microalgae and Ammonia: A Review on Inter-Relationship. Fuel 2021, 303, 121303. [Google Scholar] [CrossRef]
  74. García, D.; Alcántara, C.; Blanco, S.; Pérez, R.; Bolado, S.; Muñoz, R. Enhanced Carbon, Nitrogen and Phosphorus Removal from Domestic Wastewater in a Novel Anoxic-Aerobic Photobioreactor Coupled with Biogas Upgrading. Chem. Eng. J. 2017, 313, 424–434. [Google Scholar] [CrossRef] [Green Version]
  75. Dhaouefi, Z.; Toledo-Cervantes, A.; García, D.; Bedoui, A.; Ghedira, K.; Chekir-Ghedira, L.; Muñoz, R. Assessing Textile Wastewater Treatment in an Anoxic-Aerobic Photobioreactor and the Potential of the Treated Water for Irrigation. Algal Res. 2018, 29, 170–178. [Google Scholar] [CrossRef]
  76. Torres-Franco, A.F.; Zuluaga, M.; Hernández-Roldán, D.; Leroy-Freitas, D.; Sepúlveda-Muñoz, C.A.; Blanco, S.; Mota, C.R.; Muñoz, R. Assessment of the Performance of an Anoxic-Aerobic Microalgal-Bacterial System Treating Digestate. Chemosphere 2021, 270, 129437. [Google Scholar] [CrossRef]
  77. García, D.; Posadas, E.; Grajeda, C.; Blanco, S.; Martínez-Páramo, S.; Acién, G.; García-Encina, P.; Bolado, S.; Muñoz, R. Comparative Evaluation of Piggery Wastewater Treatment in Algal-Bacterial Photobioreactors under Indoor and Outdoor Conditions. Bioresour. Technol. 2017, 245, 483–490. [Google Scholar] [CrossRef]
  78. Nguyen, M.T.; Nguyen, T.P.; Pham, T.H.; Duong, T.T.; Do, M.V.; Trinh, T.V.; Nguyen, Q.T.X.; Trinh, V.M. Removal of Nutrients and COD in Wastewater from Vietnamese Piggery Farm by the Culture of Chlorella vulgaris in a Pilot-Scaled Membrane Photobioreactor. Water 2022, 14, 3645. [Google Scholar] [CrossRef]
  79. Zittelli, G.C.; Silva Benavides, A.M.; Silovic, T.; Ranglová, K.; Masojídek, J.; Cicchi, B.; Faraloni, C.; Touloupakis, E.; Torzillo, G. Productivity and Nutrient Removal by the Microalga Chlamydopodium fusiforme Grown Outdoors in BG-11 and Piggery Wastewater. Front. Mar. Sci. 2022, 9, 2406. [Google Scholar] [CrossRef]
  80. Madigan, M.T.; Jung, D.O.; Woese, C.R.; Achenbach, L.A. Rhodoferax antarcticus sp. Nov., a Moderately Psychrophilic Purple Nonsulfur Bacterium Isolated from an Antarctic Microbial Mat. Arch. Microbiol. 2000, 173, 269–277. [Google Scholar] [CrossRef] [PubMed]
  81. Capson-Tojo, G.; Lin, S.; Batstone, D.J.; Hülsen, T. Purple Phototrophic Bacteria Are Outcompeted by Aerobic Heterotrophs in the Presence of Oxygen. Water Res. 2021, 194, 116941. [Google Scholar] [CrossRef] [PubMed]
  82. Siefert, E.; Irgens, R.L.; Pfennig, N. Phototrophic Purple and Green Bacteria in a Sewage Treatment Plant. Appl. Environ. Microbiol. 1978, 35, 38–44. [Google Scholar] [CrossRef] [Green Version]
  83. Manhaeghe, D.; Michels, S.; Rousseau, D.P.L.; Van Hulle, S.W.H. A Semi-Mechanistic Model Describing the Influence of Light and Temperature on the Respiration and Photosynthetic Growth of Chlorella vulgaris. Bioresour. Technol. 2019, 274, 361–370. [Google Scholar] [CrossRef]
  84. Rossi, S.; Sforza, E.; Pastore, M.; Bellucci, M.; Casagli, F.; Marazzi, F.; Ficara, E. Photo-Respirometry to Shed Light on Microalgae-Bacteria Consortia—A Review. Rev. Environ. Sci. Biotechnol. 2020, 19, 43–72. [Google Scholar] [CrossRef]
  85. Mainardis, M.; Buttazzoni, M.; Cottes, M.; Moretti, A.; Goi, D. Respirometry Tests in Wastewater Treatment: Why and How? A Critical Review. Sci. Total Environ. 2021, 793, 148607. [Google Scholar] [CrossRef]
  86. Carvalho, A.P.; Meireles, L.A.; Malcata, F.X. Microalgal Reactors: A Review of Enclosed System Designs and Performances. Biotechnol. Prog. 2006, 22, 1490–1506. [Google Scholar] [CrossRef]
  87. Mata, T.M.; Martins, A.A.; Caetano, N.S. Microalgae for Biodiesel Production and Other Applications: A Review. Renew. Sustain. Energy Rev. 2010, 14, 217–232. [Google Scholar] [CrossRef] [Green Version]
  88. Marín, D.; Carmona-Martínez, A.A.; Blanco, S.; Lebrero, R.; Muñoz, R. Innovative Operational Strategies in Photosynthetic Biogas Upgrading in an Outdoors Pilot Scale Algal-Bacterial Photobioreactor. Chemosphere 2021, 264, 128470. [Google Scholar] [CrossRef] [PubMed]
  89. Marín, D.; Posadas, E.; Cano, P.; Pérez, V.; Blanco, S.; Lebrero, R.; Muñoz, R. Seasonal Variation of Biogas Upgrading Coupled with Digestate Treatment in an Outdoors Pilot Scale Algal-Bacterial Photobioreactor. Bioresour. Technol. 2018, 263, 58–66. [Google Scholar] [CrossRef] [PubMed]
  90. Alloul, A.; Cerruti, M.; Adamczyk, D.; Weissbrodt, D.G.; Vlaeminck, S.E. Operational Strategies to Selectively Produce Purple Bacteria for Microbial Protein in Raceway Reactors. Environ. Sci. Technol. 2021, 55, 8278–8286. [Google Scholar] [CrossRef]
  91. Sakarika, M.; Spanoghe, J.; Sui, Y.; Wambacq, E.; Grunert, O.; Haesaert, G.; Spiller, M.; Vlaeminck, S.E. Purple Non-sulphur Bacteria and Plant Production: Benefits for Fertilization, Stress Resistance and the Environment. Microb. Biotechnol. 2020, 13, 1336–1365. [Google Scholar] [CrossRef] [Green Version]
  92. Davis, R.; Aden, A.; Pienkos, P.T. Techno-Economic Analysis of Autotrophic Microalgae for Fuel Production. Appl. Energy 2011, 88, 3524–3531. [Google Scholar] [CrossRef]
  93. Hülsen, T.; Barry, E.M.; Lu, Y.; Puyol, D.; Keller, J.; Batstone, D.J. Domestic Wastewater Treatment with Purple Phototrophic Bacteria Using a Novel Continuous Photo Anaerobic Membrane Bioreactor. Water Res. 2016, 100, 486–495. [Google Scholar] [CrossRef] [Green Version]
  94. Chitapornpan, S.; Chiemchaisri, C.; Chiemchaisri, W.; Honda, R.; Yamamoto, K. Organic Carbon Recovery and Photosynthetic Bacteria Population in an Anaerobic Membrane Photo-Bioreactor Treating Food Processing Wastewater. Bioresour. Technol. 2013, 141, 65–74. [Google Scholar] [CrossRef]
  95. Brennan, L.; Owende, P. Biofuels from Microalgae-A Review of Technologies for Production, Processing, and Extractions of Biofuels and Co-Products. Renew. Sustain. Energy Rev. 2010, 14, 557–577. [Google Scholar] [CrossRef]
  96. Grassino, M.; Batstone, D.J.; Yong, K.W.L.; Capson-Tojo, G.; Hülsen, T. Method Development for PPB Culture Screening, Pigment Analysis with UPLC-UV-HRMS vs. Spectrophotometric Methods, and Spectral Decomposition-Based Analysis. Talanta 2022, 246, 123490. [Google Scholar] [CrossRef]
  97. Borowitzka, M.A. High-Value Products from Microalgae-Their Development and Commercialisation. J. Appl. Phycol. 2013, 25, 743–756. [Google Scholar] [CrossRef]
  98. Lorenz, R.T.; Cysewski, G.R. Commercial Potential for Haematococcus Microalgae as a Natural Source of Astaxanthin. Trends Biotechnol. 2000, 18, 160–167. [Google Scholar] [CrossRef] [PubMed]
  99. Wang, X.; Zhu, M.; Wang, K.; He, S.; Shi, X.; Yuan, B.; Dong, B.; Wang, Z. Preparation of Core-Shell Microcapsules Based on Microfluidic Technology for the Encapsulation, Protection and Controlled Delivery of Phycocyanin. J. Drug Deliv. Sci. Technol. 2022, 72, 103361. [Google Scholar] [CrossRef]
  100. Cao, K.; Zhi, R.; Zhang, G. Photosynthetic Bacteria Wastewater Treatment with the Production of Value-Added Products: A Review. Bioresour. Technol. 2020, 299, 122648. [Google Scholar] [CrossRef] [PubMed]
  101. Spolaore, P.; Joannis-Cassan, C.; Duran, E.; Isambert, A. Commercial Applications of Microalgae. J. Biosci. Bioeng. 2006, 101, 87–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Mohammadi, M.; Soltanzadeh, M.; Ebrahimi, A.R.; Hamishehkar, H. Spirulina platensis Protein Hydrolysates: Techno-Functional, Nutritional and Antioxidant Properties. Algal Res. 2022, 65, 102739. [Google Scholar] [CrossRef]
  103. Lupatini, A.L.; Colla, L.M.; Canan, C.; Colla, E. Potential Application of Microalga Spirulina platensis as a Protein Source. J. Sci. Food Agric. 2017, 97, 724–732. [Google Scholar] [CrossRef]
  104. Hülsen, T.; Hsieh, K.; Lu, Y.; Tait, S.; Batstone, D.J. Simultaneous Treatment and Single Cell Protein Production from Agri-Industrial Wastewaters Using Purple Phototrophic Bacteria or Microalgae—A Comparison. Bioresour. Technol. 2018, 254, 214–223. [Google Scholar] [CrossRef] [Green Version]
  105. Zhu, Y.; Lu, W.; Ye, L.; Chen, Z.; Hu, W.; Wang, C.; Chen, J.; Yu, H. Enhanced Synthesis of Coenzyme Q10 by Reducing the Competitive Production of Carotenoids in Rhodobacter sphaeroides. Biochem. Eng. J. 2017, 125, 50–55. [Google Scholar] [CrossRef]
  106. Kumar, A.; Kaur, H.; Devi, P.; Mohan, V. Role of Coenzyme Q10 (CoQ10) in Cardiac Disease, Hypertension and Meniere-like Syndrome. Pharmacol. Ther. 2009, 124, 259–268. [Google Scholar] [CrossRef]
  107. He, S.; Lu, H.; Zhang, G.; Ren, Z. Production of Coenzyme Q10 by Purple Non-Sulfur Bacteria: Current Development and Future Prospect. J. Clean. Prod. 2021, 307, 127326. [Google Scholar] [CrossRef]
  108. Fradinho, J.C.; Oehmen, A.; Reis, M.A.M. Improving Polyhydroxyalkanoates Production in Phototrophic Mixed Cultures by Optimizing Accumulator Reactor Operating Conditions. Int. J. Biol. Macromol. 2019, 126, 1085–1092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Chen, G.Q.; Patel, M.K. Plastics Derived from Biological Sources: Present and Future: A Technical and Environmental Review. Chem. Rev. 2012, 112, 2082–2099. [Google Scholar] [CrossRef] [PubMed]
  110. Costa, S.S.; Miranda, A.L.; de Morais, M.G.; Costa, J.A.V.; Druzian, J.I. Microalgae as Source of Polyhydroxyalkanoates (PHAs)—A Review. Int. J. Biol. Macromol. 2019, 131, 536–547. [Google Scholar] [CrossRef] [PubMed]
  111. Adessi, A.; McKinlay, J.B.; Harwood, C.S.; De Philippis, R. A Rhodopseudomonas palustris NifA* Mutant Produces H2 from NH4+ -Containing Vegetable Wastes. Int. J. Hydrogen Energy 2012, 37, 15893–15900. [Google Scholar] [CrossRef]
  112. Fan, L.; Tu, Z.; Chan, S.H. Recent Development of Hydrogen and Fuel Cell Technologies: A Review. Energy Rep. 2021, 7, 8421–8446. [Google Scholar] [CrossRef]
  113. Adessi, A.; De Philippis, R. Photobioreactor Design and Illumination Systems for H2 Production with Anoxygenic Photosynthetic Bacteria: A Review. Int. J. Hydrog. Energy 2014, 39, 3127–3141. [Google Scholar] [CrossRef]
  114. Iqbal, K.; Saxena, A.; Pande, P.; Tiwari, A.; Chandra Joshi, N.; Varma, A.; Mishra, A. Microalgae-Bacterial Granular Consortium: Striding towards Sustainable Production of Biohydrogen Coupled with Wastewater Treatment. Bioresour. Technol. 2022, 354, 127203. [Google Scholar] [CrossRef]
  115. Arutselvan, C.; Seenivasan, H.K.; Lewis Oscar, F.; Ramya, G.; Thuy Lan Chi, N.; Pugazhendhi, A.; Thajuddin, N. Review on Wastewater Treatment by Microalgae in Different Cultivation Systems and Its Importance in Biodiesel Production. Fuel 2022, 324, 124623. [Google Scholar] [CrossRef]
  116. Szulczyk, K.R.; Tan, Y.M. Economic Feasibility and Sustainability of Commercial Bioethanol from Microalgal Biomass: The Case of Malaysia. Energy 2022, 253, 124151. [Google Scholar] [CrossRef]
  117. Ho, S.H.; Huang, S.W.; Chen, C.Y.; Hasunuma, T.; Kondo, A.; Chang, J.S. Bioethanol Production Using Carbohydrate-Rich Microalgae Biomass as Feedstock. Bioresour. Technol. 2013, 135, 191–198. [Google Scholar] [CrossRef] [PubMed]
  118. Acebu, P.I.G.; de Luna, M.D.G.; Chen, C.Y.; Abarca, R.R.M.; Chen, J.H.; Chang, J.S. Bioethanol Production from Chlorella vulgaris ESP-31 Grown in Unsterilized Swine Wastewater. Bioresour. Technol. 2022, 352, 127086. [Google Scholar] [CrossRef] [PubMed]
  119. Appels, L.; Baeyens, J.; Degrève, J.; Dewil, R. Principles and Potential of the Anaerobic Digestion of Waste-Activated Sludge. Prog. Energy Combust. Sci. 2008, 34, 755–781. [Google Scholar] [CrossRef]
  120. Muñoz, R.; Meier, L.; Diaz, I.; Jeison, D. A Review on the State-of-the-Art of Physical/Chemical and Biological Technologies for Biogas Upgrading. Rev. Environ. Sci. Biotechnol. 2015, 14, 727–759. [Google Scholar] [CrossRef] [Green Version]
  121. Xu, J.; Zhao, Y.; Zhao, G.; Zhang, H. Nutrient Removal and Biogas Upgrading by Integrating Freshwater Algae Cultivation with Piggery Anaerobic Digestate Liquid Treatment. Appl. Microbiol. Biotechnol. 2015, 99, 6493–6501. [Google Scholar] [CrossRef] [PubMed]
  122. Zabed, H.M.; Akter, S.; Yun, J.; Zhang, G.; Zhang, Y.; Qi, X. Biogas from Microalgae: Technologies, Challenges and Opportunities. Renew. Sustain. Energy Rev. 2020, 117, 109503. [Google Scholar] [CrossRef]
  123. Batool, K.; tuz Zahra, F.; Rehman, Y. Arsenic-Redox Transformation and Plant Growth Promotion by Purple Nonsulfur Bacteria Rhodopseudomonas palustris CS2 and Rhodopseudomonas faecalis SS5. Biomed. Res. Int. 2017, 2017, 6250327. [Google Scholar] [CrossRef] [Green Version]
  124. Coppens, J.; Grunert, O.; Van Den Hende, S.; Vanhoutte, I.; Boon, N.; Haesaert, G.; De Gelder, L. The Use of Microalgae as a High-Value Organic Slow-Release Fertilizer Results in Tomatoes with Increased Carotenoid and Sugar Levels. J. Appl. Phycol. 2016, 28, 2367–2377. [Google Scholar] [CrossRef]
Figure 1. Diagram of photosynthetic microorganisms classification performing anoxygenic or oxygenic photosynthesis.
Figure 1. Diagram of photosynthetic microorganisms classification performing anoxygenic or oxygenic photosynthesis.
Symmetry 15 00525 g001
Figure 2. Metabolic diagram of purple phototrophic bacteria, based on the genome of Rhodopseudomonas palustris (adapted from [46]).
Figure 2. Metabolic diagram of purple phototrophic bacteria, based on the genome of Rhodopseudomonas palustris (adapted from [46]).
Symmetry 15 00525 g002
Figure 3. Diagram of electron transport (black lines) in anoxygenic (A) and oxygenic (B) photosynthesis in photosynthetic microorganisms. Light-harvesting (LH), bacteriochlorophyll (Bph), quinone (Q), cytochrome C2 (C2), photosystem (PS), plastoquinone (PQ), plastocyanin (PC), chlorophyll (Chl), ferredoxin (Fd), NADP oxidoreductase (FNR) (adapted from [39]).
Figure 3. Diagram of electron transport (black lines) in anoxygenic (A) and oxygenic (B) photosynthesis in photosynthetic microorganisms. Light-harvesting (LH), bacteriochlorophyll (Bph), quinone (Q), cytochrome C2 (C2), photosystem (PS), plastoquinone (PQ), plastocyanin (PC), chlorophyll (Chl), ferredoxin (Fd), NADP oxidoreductase (FNR) (adapted from [39]).
Symmetry 15 00525 g003
Figure 4. Absorption spectrum (400–1150 nm) of the purple phototrophic bacteria (A) Rhodobacter capsulatus (1), Rhodobacter sphaeroides (2), Rhodospirillum rubrum (3), Roseospirillum parvum (4), Blastochloris viridis (5), the microalgae (B) Phaeodactylum tricornutum (6), Palmaria palmata (7), Isochrysis sp. (8), Chlamydomonas sp. (9), the green sulphur bacteria (C) Prosthecochloris aestuarii (10), Pelodictyon phaeoclathratiforme (11) and the cyanobacteria (D) Synechococcus WH7803 (12), Synechococcus WH8103 (13), Synechococcus BS5 (14), Synechococcus BS4 (15), Prochlorococcus sp. (16) and Acaryochloris marina (17) (adapted from [48]).
Figure 4. Absorption spectrum (400–1150 nm) of the purple phototrophic bacteria (A) Rhodobacter capsulatus (1), Rhodobacter sphaeroides (2), Rhodospirillum rubrum (3), Roseospirillum parvum (4), Blastochloris viridis (5), the microalgae (B) Phaeodactylum tricornutum (6), Palmaria palmata (7), Isochrysis sp. (8), Chlamydomonas sp. (9), the green sulphur bacteria (C) Prosthecochloris aestuarii (10), Pelodictyon phaeoclathratiforme (11) and the cyanobacteria (D) Synechococcus WH7803 (12), Synechococcus WH8103 (13), Synechococcus BS5 (14), Synechococcus BS4 (15), Prochlorococcus sp. (16) and Acaryochloris marina (17) (adapted from [48]).
Symmetry 15 00525 g004
Figure 5. Metabolic interactions between microalgae and bacteria during wastewater treatment.
Figure 5. Metabolic interactions between microalgae and bacteria during wastewater treatment.
Symmetry 15 00525 g005
Figure 6. Schematic diagram of common photobioreactors for photosynthetic microorganisms cultivation. Natural lake (A), open pond (B), open raceways (C), vertical flat-panel (D), closed tubular (E) and photo-anaerobic membrane bioreactor (F).
Figure 6. Schematic diagram of common photobioreactors for photosynthetic microorganisms cultivation. Natural lake (A), open pond (B), open raceways (C), vertical flat-panel (D), closed tubular (E) and photo-anaerobic membrane bioreactor (F).
Symmetry 15 00525 g006
Figure 7. The electromagnetic spectrum, extension of visible spectrum (400–780 nm) and infrared spectrum (780–1,000,000 nm). Infrared spectrum is composed of near-infrared (NIR: 0.780–3000 nm), mid-infrared (MIR: 3–50 µm) and far-infrared (FIR: 50–1000 µm).
Figure 7. The electromagnetic spectrum, extension of visible spectrum (400–780 nm) and infrared spectrum (780–1,000,000 nm). Infrared spectrum is composed of near-infrared (NIR: 0.780–3000 nm), mid-infrared (MIR: 3–50 µm) and far-infrared (FIR: 50–1000 µm).
Symmetry 15 00525 g007
Figure 8. Summary of different potential products with high added value in photosynthetic microorganisms (PPB and microalgae). Gray frames require auxiliary processes.
Figure 8. Summary of different potential products with high added value in photosynthetic microorganisms (PPB and microalgae). Gray frames require auxiliary processes.
Symmetry 15 00525 g008
Table 1. Characteristic composition in domestic, industrial an agro-industrial wastewater.
Table 1. Characteristic composition in domestic, industrial an agro-industrial wastewater.
CompositionDomesticIndustrial 1Agro-Industrial 2
COD (mg L−1)52610,00054,000
TN (mg L−1)4615000
TP (mg L−1)631500
pH6.87.07.7
1 Industrial wastewater represented by brewery wastewater [7]. 2 Agro-industrial wastewater represented by piggery wastewater [8].
Table 3. Summary of piggery wastewater treatment using purple phototrophic bacteria.
Table 3. Summary of piggery wastewater treatment using purple phototrophic bacteria.
PWW Characteristics
(mg L−1)
Reactor Type
(Volume)
Dominant StrainPollutant Removal (%)References
CarbonNitrogen
TOC: 574Open PBR Rhodoplanes sp.8783[30]
TN: 166(3 L)
TOC: 10,318–1989Batch PBR PPB mixed7813[31]
TN: 2209–563(0.4 L)
TOC: 15,775–1131Batch PBR Rhodopseudomonas sp.7539[59]
TN: 5028–366(0.5 L)
TOC: 1180Open PBR PPB mixed9999[14]
TN: 380(3 and 1.5 L)
COD: 4130Flat plate PBR Rhodopseudomonas sp.7122[3]
TKN: 1160(60, 80 and 100 L)
TOC: 860Batch PBR R. palustris7942[9]
TN: 380(0.5 L)
Table 4. Summary of piggery wastewater treatment using microalgae.
Table 4. Summary of piggery wastewater treatment using microalgae.
PWW Characteristics
(mg L−1)
Reactor Type
(Volume)
Dominant StrainPollutant Removal (%)References
CarbonNitrogen
COD: 526–4346High-rate algal ponds Microalgae mixed7688[8]
TKN: 59–370(464 L)
TOC: 963Open PBR C. vulgaris9456[77]
TN: 341(3 L)
TOC: 574Open PBR Microalgae mixed8487[30]
TN: 166(3 L)
COD: 8420–6870Sequential Batch PBR Microalgae mixed9290[21]
NH3-N: 850–1150(4 L)
COD: 287Membrane PBR C. vulgaris6674[78]
NH4+: 184(50 L)
COD: 1016Cylindrica PBR Chlamydopodium
fusiforme
4193[79]
N-NH4: 92(10 L)
Table 5. Principal advantages and disadvantages of open and closed configurations of photobioreactors used for the cultivation of photosynthetic microorganisms.
Table 5. Principal advantages and disadvantages of open and closed configurations of photobioreactors used for the cultivation of photosynthetic microorganisms.
Open
Photobioreactors
Closed
Photobioreactors
Reference
Capital investmentLowHigh[91]
ScalabilityHighVariable[92]
Culture controlLowHigh[64]
Culture contaminationHighLow[64]
Evaporation rateHighLow[64]
Biomass productivityVariableHigh[92]
Nutrient removalHighHigh[65]
Nutrient recovery in biomassVariableHigh[92]
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

Sepúlveda-Muñoz, C.A.; de Godos, I.; Muñoz, R. Wastewater Treatment Using Photosynthetic Microorganisms. Symmetry 2023, 15, 525. https://doi.org/10.3390/sym15020525

AMA Style

Sepúlveda-Muñoz CA, de Godos I, Muñoz R. Wastewater Treatment Using Photosynthetic Microorganisms. Symmetry. 2023; 15(2):525. https://doi.org/10.3390/sym15020525

Chicago/Turabian Style

Sepúlveda-Muñoz, Cristian A., Ignacio de Godos, and Raúl Muñoz. 2023. "Wastewater Treatment Using Photosynthetic Microorganisms" Symmetry 15, no. 2: 525. https://doi.org/10.3390/sym15020525

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