Open access peer-reviewed chapter

Anaerobic Co-Digestion of Microalgae and Industrial Wastes: A Critical and Bibliometric Review

Written By

David de la Lama-Calvente, Juan Cubero, María José Fernández-Rodríguez, Antonia Jiménez-Rodríguez and Rafael Borja

Submitted: 21 February 2022 Reviewed: 07 March 2022 Published: 19 May 2022

DOI: 10.5772/intechopen.104378

From the Edited Volume

Progress in Microalgae Research - A Path for Shaping Sustainable Futures

Edited by Leila Queiroz Zepka, Eduardo Jacob-Lopes and Mariany Costa Deprá

Chapter metrics overview

149 Chapter Downloads

View Full Metrics

Abstract

Microalgae are photosynthetic organisms able to grow faster than land plants and produce biomass with relatively high energy potential. Accumulated high-value compounds like lipids, minerals, or proteins have focused the attention of scientists due to the potential production of biofuels and other value-added products. However, several drawbacks regarding both the biochemical structure of these organisms and technological difficulties have prevented the industry for implementing a comprehensive low-cost process regarding energy and environmental contamination. Among these technologies, anaerobic digestion (AD) has greatly increased research attention because of its simplicity and the ability to produce easily recycle by-products. Moreover, anaerobic co-digestion (AcoD) has shown promising results as a method to bypass the AD problems of microalgae as a sole substrate. This review is focused on the recent trends and comparison of the AcoD process to maximize energy recovery from microalgae biomass and agro-industrial wastes. The yield of methane gas among the studied bibliography is compared and a critical review of published data and methods used is included.

Keywords

  • anaerobic co-digestion
  • microalgae
  • methane production
  • review methodology
  • agro-industrial wastes

1. Introduction

Microalgae are a wide family of photosynthetic organisms able to increase their biomass by using CO2 and sunlight as energy sources by a rate 100 times faster than plants [1]. Moreover, the water and nutrients consumption is lesser than the needed for the same amount of biomass of terrestrial crops and it does not compete with other biomass from land areas [1, 2].

For the above-mentioned reasons, microalgae have been studied for decades for their potential conversion to energy. However, it was not until recently that microalgae have been considered feasible biomass to be used as a feedstock to produce biofuels (e.g. biohydrogen, biodiesel, biomethane, bio-oil, bioethanol, etc.) [1, 2, 3, 4]. This is mainly due to several drawbacks (e.g. the need of solvents to produce biodiesel contributing to greenhouse gas emission, the use of expensive enzymes for bioethanol production, etc.) that have been overpassed thanks to technology and research efforts over the years focusing on the understanding and optimizing those factors that affect the different systems along with a better understanding of the biomass itself (e.g. the effect of the cell wall, the algae growth requirements, etc.) [1, 2, 3, 4, 5].

Among these technologies, anaerobic digestion (AD) has shown promising results. AD is a biological process where the organic compounds from certain biomass are degraded in the absence of oxygen (O2) by a microbial consortium. The main effluents of this process are biogas (i.e. a gas composed primarily of methane and CO2) and a nutrient-rich digestate [3]. While the produced biogas is considered a renewable energy source to produce electricity and heat in cogeneration plants, the nutrient-rich digestate could be used as a fertilizer [3, 4].

Microalgae are not only a viable AD feedstock, but can also serve as a means of biogas upgrading and its cultivation in the digestate can reduce the excess of nutrients and mitigate its potential toxicity [5]. However, the mono-digestion of microalgae has shown some concerns regarding its viability at industrial scales. Briefly, these concerns are related to the presence of long-chain organic compounds, mainly in the cell wall, the low carbon to nitrogen (C/N) ratio, and the high retention time needed in the reactors, which led to low methane yields, undigested organic matter in the digestate and more importantly to the inhibition of the AD process [2, 3].

In order to overpass these problems, anaerobic co-digestion (AcoD) of microalgae along with a wide range of co-substrates has been the focus of several research groups recently [1, 2, 3, 4, 5, 6, 7]. AcoD has several benefits due to its capacity to enhance the C/N ratio, the buffer capacity, the nutrient balance, and to dilute inhibitory compounds [4]. These improvements produce higher methane yields which in most cases are higher than the theoretical values obtained from the sole digestion of each co-substrate showing a synergetic effect [1, 2, 3, 4, 5].

Nevertheless, AcoD presents some drawbacks such as a higher organic load in the digestate, the usual need of pretreatments, or the difficulties to maintain a stable feedstock along the seasons [3]. This chapter aims to summarize the state of the art of microalgae used as co-substrate in AcoD processes and to highlight knowledge gaps and potential future developments. Moreover, a comprehensive analysis of the parameters affecting AcoD of microalgae in order to enhance methane yield is included in detail. Lastly, the energetic viability of several scenarios is discussed and future trends proposed.

Advertisement

2. Review methodology

For this bibliographic review, 92 articles have been selected from the 137 articles were carried out on the Scopus with the keywords “microalgae” and “anaerobic co-digestion” during the periods of 2011–2021 [8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55, 56, 57, 58, 59, 60, 61, 62, 63, 64, 65, 66, 67, 68, 69, 70, 71, 72, 73, 74, 75, 76, 77, 78, 79, 80, 81, 82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97, 98, 99, 100]. The bibliometric analysis was realized using the VOSviewer software using these 92-original articles.

Thanks to this software, it has been verified that there are few groups dedicated to the study of AcoD with microalgae. These investigations focus mainly on countries such as Spain (30 articles), the United States (17 articles), China (8 articles), Mexico (8 articles), and Brazil (5 articles). In addition, only eight authors have five or more articles to 321 authors (Figure 1a). In Figure 1b, as can see the connection between different authors and groups. The color and size indicate the group and the citations of these eight authors.

Figure 1.

Bibliometric analysis with VOSviewer: (a) authors with five or more articles, (b) connexion between authors, (c) substrates and (d) microalgae.

Advertisement

3. Microalgae and growth mediums

Since the first study on biogas production with microalgae by Golueke [101], a wide variety of microalgae genera have been studied. Due to the composition of the microalgae, it has been seen that each strain has a very specific biogas production, and has very diverse productions. One of the factors that most affect the AD of microalgae is the structure of its cell wall, which is why the selection of the type of microalgae is important. One of the main characteristics for microalgae to have a good methane production potential is to have a thin or null cell wall, large cytoplasmic components, a high growth rate, and a high tolerance to stress [102]. Other important aspects are whether it has a low content of hollocellulose in the cell walls, metabolic, and growth conditions are favorable, the morphological traits of the microalgae strains [27]. In addition, the selected strains offer a feasible genetic manipulation to control metabolic activities and improve tolerance to nutrient and ecological stress [102].

Nineteen genera of microorganisms have been investigated in the 92 articles found on the AcoD of microalgae. Of these 92 articles, 51 of these works have been studied with the genus Chlorella and 24 of these works with the genus Scenedesmus. The rest of the work carried out was with Chlorophyta as Nannochloropsis (5 studies), Micractinium (3 studies), Dunaliella (2 studies), Dictyosphaerium, Closteriopsis, Desmodesmus, Chlamydomonas, Stigeoclonium, Botryococcus and Tetraselmis genus. Other genus as Tribonema (Ochrophyta), Phaeodactylum (Bacillariophyta) and Tisochrysis (Haptophyta) have been also studied as co-substrate. In addition, the following cyanobacteria have been studied: Arthrospira (4 studies), Spirulina (2 studies), Merismopedia, Oscillatoria.

Another factor that should be considered is the culture medium for microalgae growth. Depending on the medium, it could favor the production of biogas in a later step due to the nutrient requirement of the microorganisms in the AD process [96]. According to the reviewed bibliography, the microalgae used for AcoD are obtained through other research groups or are cultivated using three different types of medium for growing it.

  1. Synthetic medium: BG11, Bold’s basal (BBM), Conway enriched medium, Jaworki’s medium, modified Zarrouk medium, BlueBIOTech Ltf. (Table 1).

  2. Digestate or effluent anaerobic: Anaerobic sludge, continuous stirred tank reactor (CSTR) digestate, up-flow anaerobic sludge blanket (UASB) digestate, primary effluent and sludge, chicken manure digestate, swine digestate, Anaerobic membrane bioreactor (AnMBR) digestate.

  3. Wastewater: Tannery effluents, piggery WW, industry, domestic, municipal, fresh waters, lake waters, natural seawater enriched, soft drink. WW, Winery WW.

As shown in Table 1, microalgae are microorganisms that need certain nutrients to perform vital functions. Therefore, synthetic culture media contain macronutrients such as calcium, sodium, potassium, magnesium, and chloride. In addition to adding micronutrients such as iron, cobalt, molybdenum, manganese, copper, zinc, and vitamins. Finally, ethylenediaminetetraacetic acid (EDTA) is also added to form a complex ring (a chelate) with the trace elements, which, when used in low concentrations, stimulates the growth of microalgae by making this element available in low quantities. These nutrients are found in nature and are bioavailable in the other two natural culture media such as anaerobic digestate and wastewater, but may be found in lower concentrations than necessary [10].

NutrientBG11BBMConwayJaworki’sZarrouk
NitrateNaNO3NaNO3KNO3Ca(NO3)2·4H2O NaNO3NaNO3
PotassiumK2HPO4K2HPO4 KH2PO4Na3PO4KH2PO4 Na2HPO4·12H2OK2HPO4 K2SO4
SodiumNa2CO2NaHCO3NaHCO3 and Na2CO3
MagnesiumMgSO4·7H2OMgSO4·7H2OMgSO4·7H2OMgSO4·7H2O
ChlorideCaCl2·2H2ONaCl CaCl2·2H2ONaCl CaCl2
BoronH3BO3H3BO3H3BO3H3BO3
ZincZnSO4·7H2OZnSO4·7H2OZnCl2
ManganeseMnCl2·4H2OMnCl2·4H2OMnCl2·4H2OMnCl2·4H2O
MolybdenumNa2MoO4·2H2OMoO3(NH4)6Mo7O24·4H2O(NH4)6Mo7O24·4H2O
CopperCuSO4·6H2OCuSO4·6H2OCuSO4·6H2O
CobaltCo(NO3)2·6H2OCo(NO3)2·6H2OCoCl2·6H2O
IronFerric ammonium citrateFeSO4·7H2OFeCl3·6H2ONaFeEDTAFeSO4·7H2O
EDTANa2EDTANa2EDTA (KOH)Na2H2EDTA·2H2ONa2EDTAEDTA
VitaminCitric acidThiamin HCl
CyanocobalaminThiamin HCl
Cyanocobalamin Biotin

Table 1.

Synthetic culture medium for the growth of microalgae.

In addition, using wastewater for the growth of microalgae could prevent eutrophication of the water due to the consumption of excess nutrient of these type of waters. In this sense, microalgae have been succesfully used for removing nitrogen and phosphorous from various wastes as shown in Table 2. Significant amounts of removal of nutrients and biomass production achieved in these studies demonstrate the feasibility of coupled wastewater treatment and microalgae cultivation processes.

AlgaeN-NH4P-PO4Ref.
[Conc] mg/L%Remove[Conc] mg/L%Remove
Scenedesmus sp. + Chlorella sp.7.9972.593[10]
Mix culture35.561nd*nd*[80]
Arthrospira platensis37588.415897.01[42]
Chlorella vulgaris28599.611791.2[13]
Oscillatoria tenius10.296.10.882.9[18]
Chlorella 1067202.6894.337.1829.67[38]
Chlorella sp.11445818898[84]
Micractinium nov2.294495[82]
Chlorella sp.1.5963.595
Micractinium nov27920.751[88]
Chlorella sp.
Spirulina platensis120685530[92]
Chlorella sp.2258512020
Cholerella sp.4588958[56]

Table 2.

Removal of N-NH4 and P-PO4 in microalgae culture.

nd: not determined.


Advertisement

4. Co-substrates

AcoD is the use of a mixture of biomasses to obtain a relatively higher methane yield [10]. This modified technique is considered economically more viable and easier to control mixed biomass compared to traditional mono-digestion systems. Depending on the species, microalgae contain significantly high or low amounts of protein, carbohydrates, and lipids. To balance the nutritional requirements of microorganisms in anaerobic reactors [13]. In addition, it would be possible to improve the stabilization of the process with well-balanced mixtures. With this, it could be possible to increase the organic load capacity, reduce the concentration of possible inhibitors, and increase the buffering capacity of the digestates. Apart from the nutrient balances and net synergistic effects that would occur when co-digesting microalgae with an efficient substrate. In the revised bibliography, synthetic co-substrates, agri-food residues and slurry, and liquid residues can be found.

  1. Synthetic substrate: Cellulose and glycerol, and synthetic food wastes.

  2. Solid waste: Cattle manure (chicken, cow, pig), cheese whey powder, chicken litter, green willow, chromium tanned leather shavings (SHA), coffee wastes, corn silage, fat, oil and grease (FOG), food waste, millet grass (Pennisetum glaucum), model kitchen waste, olive mill solid waste (OMSW), oil palm empty fruit bunches, Opuntia maxima, papaya waste, potato wastes, rice straw, Sida hermaphrodita (L.), silage and seaweed, sugarcane leaves, teak leaves (Tectona grandis), used cooking oil.

  3. Wastewater: Bacteria biomass from anaerobic sludge, biosolids of water resource recovery facilities, catering waste leachate and raw sludge, cow rumen liquid, deproteinated cheese whey, dewatered ww, food waste leachate, mill residue, municipal, paper sludge, palm oil mill effluent, septic tank sludge (STS), sewage sludge, swine wastewaters, waste activated sludge (WAS), and thickened WAS.

The production and processing of food and feed results in the generation of a large amount of waste. AD stands out as a suitable technology to reduce the environmental impact of agro-industrial waste and increase the energy self-sufficiency of these industries [29]. However, agro-waste is characterized by the lack of nutrients in its composition necessary for AD. In addition, having a high C/N ratio can affect the performance of AD. Another waste used as a co-substrate is animal manure (i.e. pig, cattle, and poultry). Contrary to agri-food residues, it has a relatively low C/N ratio, which increases the risk of ammonia inhibition [102]. Finally, one of the most studied co-substrates is sewage sludge, which was studied for the first time in 1983 [103], when it was co-digested with Spirulina maxima. Sludge of various varieties can be used as co-substrate, although like the rest, the interactions between microalgae-substrate must be studied to obtain a good balance between the different parameters that can affect AD [12].

Advertisement

5. Factors influencing anaerobic co-digestion

AD process is affected by several factors which led to a higher or lower biogas yield; these factors can be split into two main sources: operational conditions and substrates composition. Although both sources are widely related, they can be studied and controlled separately. The main operational conditions are temperature, pH, configuration of bioreactors, acclimation of inoculum, hydraulic retention time (HRT), organic load rate (OLR), and inoculum to substrate ratio (ISR) [1, 104]. The main factors related to substrates composition are the C/N ratio and macro and micronutrients [1, 5, 6, 104].

In this sense, AcoD is proposed as a feasible alternative in order to balance these factors and allow a better performance and a higher biogas production.

5.1 Effect of initial conditions

Most studies agree that the best operational temperatures for AD are under mesophilic (20–45°C) or thermophilic (>45°C) conditions [104]. Temperature affects, either directly or indirectly, the solubility of substrate compounds and the specific growth rate of the microorganisms involved, provoking a change in the HRT, the pH, and the methane yield [104]. Among the literature, the most common range is the mesophilic conditions due to its lower energy cost requirements and the similar methane yield when compared with higher temperature conditions [105]. However, temperature variations during AD performance have shown significant reductions in methane yield and the kinetics of the process [104, 106].

The effect of pH is mainly related to the optimum pH of the microorganism performance during AD. Based on that, a pH between 4.6 and 6.0 favored the hydrolysis, acidogenesis, and acetogenesis stages, while a pH between 6.0 and 8.0 favored the methanogenesis phase [1]. Literature shows that the most suitable pH range is between 6.5 and 7.5, where the methane production is most benefited [1, 104]. Furthermore, it has been reported that the initial pH of the substrate had a significant impact on methane yield, being the optimum value in the range of 7.0–7.5 for the co-digestion of swine manure and maize stalk [107]. However, pH is highly affected during the AD process, lowering its value if the buffer system is not strong enough due to the accumulation of volatile fatty acids (VFA) (e.g. acetic, propionic, butyric, and valeric acids) [104] or increasing it if ammonium nitrogen is accumulated (around 5.0 g NH3-N L−1) [33].

Temperature and pH have been linked to free ammonium nitrogen and ammonium ions equilibrium, showing that when one parameter is fixed there is a linear increase in methane yield when the other two independent variables increased up to a certain limit, after which the methane production decrease [104, 108]. A recent study showed optimal conditions for the mono-digestion of chicken manure of 34.0°C, 5.0 g NH3-N L−1, and pH 7.5 [108]. Moreover, in an earlier study, an increase of pH from 7 to 8 and above, enhanced the biogas production with similar methane proportion when temperature conditions increased from 37 to 55°C during the AD of buffalo manure [109].

As shown above, temperature and pH have a significant impact on AD performance and biogas production. Thus, co-digestion is presented as a suitable technique able to enhance the buffer system, the substrate pH, and free ammonium nitrogen values, allowing a higher methane yield and a more stable process [67, 77, 78, 99, 100]. Meneses-Reyes et al. showed that increasing the C/N ratio of the substrate by reducing the microalgae ratio provoked an increase in pH, however, the pH of digestate is similar among the different co-digestion ratios studied (7.33–7.51) [41]. AcoD of Chlorella sp. and food waste in batch mode (35°C) showed that methane yield was related to the initial pH of the substrates (7.3–8.7), being the optimum value 8.0 and reporting that the methane yield decreased almost linearly when pH differs from the optimum value, although these results were not conclusive as other variables were also different (e.g. VSfeeding from 8.0 to 9.2 g) [35]. However, a mixture of microalgae biomass with thermally treated wheat straw presented a pH of 12 with no significant effect on AcoD performance when compared with the AcoD of microalgae biomass and untreated wheat straw (pH: 6.82) [71]. Another study assessed the effect of temperature within the mesophilic range (25°C, 30°C, 35°C, 40°C) in biogas production, reporting a significant increase of biogas production (45%) when the temperature was increased up to 35°C, but a reduction of production or no significant improvement (depending on the C/N ratio of substrates) when the temperature was set up at 40°C [24]. This result is in accordance with other studies reporting a decrease of methane yield for the AcoD of microalgae with undigested sewage sludge in batch mode when increasing the temperature from mesophilic (37°C) to thermophilic (55°C) conditions, even though the pH was not affected (6.91–7.03) and the processes were stable [45]. AcoD of corn silage and Nannochloropsis salina in a semi-continuous mode (38°C; C/N: 21.2) do not present any stability deviation by the increasing OLR (2–4.7 g VS L−1 d−1), while the AcoD of two microalgae (Scenedesmus sp. and Opuntia maxima) in semi-continuous mode (37°C; C/N: 15.6) showed that pH was affected by OLR (2–6.67 g VS L−1 d−1) and ISR (6:8% VS basis), is the most stable conditions OLR: 2 g VS L−1 d−1 and ISR: 8% VS basis [60, 110]. Furthermore, a study assessing the effect of alkali, acid, and thermal pretreatment of Oscillatoria tenuis, before the AcoD with pig manure, reported that pH control affected the biogas production rather than physical or chemical pretreatments [18]. Similar results were reported during the AcoD of alkali pretreated microalgae consortium with swine wastewater, where the negative effect of ammonia inhibition at high pH (11) was stronger than the positive effect of the destruction of microalgae’s cell walls [47].

pH is also being widely used as a control parameter able to indicate the stability of the reactors since is strongly linked to VFA accumulation [40, 43]. AcoD of Arthrospira platensis with several carbon-rich co-substrates proved to be stable at low OLR (1 g VS L−1 d−1) in a semi-continuous mode but unstable at higher OLR (2–5 g VS L−1 d−1) due to VFA accumulation and pH dropped, this study is in accordance with the AcoD of Chlorella sp. and glycerol that presented a stable pH range (6.6–7.32) when the HRT was above 5 days, at which point volatile fatty acids (VFA) accumulation inhibit the biogas production [24, 33]. However, AcoD of naturally grown microalgae consortium with WAS in a semi-continuous mode (37°C) showed that when the HRT was increased from 1 to 3 to 4–6, pH was reduced (from 7.51 to 7.04), although VFA accumulation was not observed and the system remained stable with no difference in methane production except at HRT of 6 were slightly dropped [46].

Another important factor is alkalinity, which helps to prevent large changes in pH due to the accumulation of volatile fatty acids, or the generation of ammonia due to protein hydrolysis. Alkalinity provides the necessary buffering capacity to counteract possible changes in pH, produced by the balance between carbonate and bicarbonate. The ideal alkalinity values for AD would be between 2000 and 4000 mg CaCO3 L−1 [99, 100]. A study assessed the relation between pH and alkalinity, where the initial pH was fixed at 7.0 while the initial alkalinity changed (70–3200 mg CaCO3 L−1), however, pH remained around neutral values (6.9–7.2) during the AcoD process suggesting that initial alkalinity has no impact avoiding an ammonium concentration from nitrogen-rich substrates as microalgae [30].

The C/N ratio is another factor that influences the AD process [29]. A good substrate C/N ratio can range between 20 and 30, with an optimal value of 25 [102]. With a C/N ratio below 20, there is an imbalance between C and N in the reactor, which ultimately releases a large amount of NH3, which usually happens with the degradation of microalgae [102]. The high concentration of NH3 in the digester affects the growth and metabolism of microorganisms and produces an accumulation of volatile fatty acids, which results in a decrease in biogas yield. This factor can also be supplied by choosing a good co-substrate and optimizing this C/N ratio [94].

The ISR is another key parameter that influences AD and methane production [13]. To find the maximum methane potential, a proper balance between the substrate and the microorganisms is necessary so that limitations and inhibitions do not occur due to the loading of the substrate. For biochemical methane potential (BMP) assays, a ISR ≥ 2 is suggested as the default value [111].

5.2 Effect of pretreatments

As a common way to improve the methane yield of AcoD of microalgae, several studies have reported the effect of pretreatments on microalgae biomass. Table 3 shows some of these pretreatments and the effect on AcoD calculated as the increase of biomethane production in percentage over the non-pretreated AcoD.

PretreatmentConditionsMicroalgaeCo-substrateImprovementReference
Enzymatic0.1% v/v; 0.5 hScenedesmus sp. + Chlorella sp.WAS-2%[97]
Ultrasonicationn/dNannochloropsis oculataCow manure16.7%[57]
Hot water121°C; 15 psiNannochloropsis oculataCow manure36.7%[57]
Ultrasonication:hot water121°C; 15 psiNannochloropsis oculataCow manure16.7%[57]
Thermal60°C, 24 hScenedesmus sp. + Chlorella sp.WAS−33.8%[86]
Thermal55°C, 2dChlorella vulgarisWW8.3%[19]
Thermal120°C, 20 minOscillatoria teniusPig manure43%[18]
AlkalineNaOH 3% w/vScenedesmus sp. + Chlorella sp.Swine WW−10.4%[47]
Thermal120°C, 1 hChlorella sp.Coffee wastes13.9%[49]
Thermo-alkaline10%CaO, 75°C, 24 hChlorella sp.Wheat straw9.0%[71]
Thermo-alkaline10%CaO, 75°C, 24 hChlorella sp.Wheat Straw15%[71].

Table 3.

Effect of pretreatment on AcoD.

It has been reported that although some pretreatments can improve the methane yield of microalgae as a sole substrate, these have a negative effect during the AcoD, mainly due to the high organic matter consumption or the inefficacy of pretreatments breaking down the cell wall [47, 86, 97].

From Table 3 it can be seen that low thermal pretreatments (60–55°C; 1−2d) have none or very little effect on AcoD [19, 86], however, when the temperature increased to 120°C, the biomethane production improve greatly (up to 43%) [18, 49]. Other successfully tested pretreatments are ultrasonication, hot water, and a combination of thermal and alkaline pretreatments [57, 71]. Nevertheless, based on the higher improvement on biomethane production, thermal pretreatment at 120°C is the most effective process, where time and pressure would be the variables to analyze [49, 57].

Advertisement

6. Biomethane potential (BMP) performance

Table 4 summarizes all the BMP assay results up to date complying with the following prerequisites:

  • The co-substrates ratio is well described as VS or C/N.

  • ISR has been considered.

  • Methane yield have been reported under standard conditions.

Co-substratesCo-substrate ratio (VS)ISR (VS)Temperature (°C)Methane yield (NLCH4kgVS-1)Synergetic effect (%)Ref.
Dictyosphaerium sp.:synthetic food wastes1:323551427.6[29]
Tribonema sp.:pig manure2:8235580.416.9[34]
Chlorella sp.:sludge:FOG4:4:22353343.7[72]
Scenedesmus quadricauda:OMSW1:3, C/N: 25.323546136.4[26]
Chlamydomonas reinhardtii 6145:OMSW1:1, C/N: 18.323554237.2[27]
Chlamydomonas reinhardtii cw15:OMSW1:1, C/N: 18.323545116.5[27]
Nannochloropsis limnetica:piggery slurry4:655335512[77]
Microalgae consortium:WW1:32353395.9[74]
Scenedesmus sp.:deproteinated cheese whey17:833.335302−7.6[16]
Scenedesmus sp.:cellulose16:843.335272−2.0[16]
Microalgae consortium:WW37:63235237.12.1[46]
Chlorella vulgaris:cattle manure4:14554318–15[40]
Chlorella sp.:swine manure6:94, C/N: 33.923534811.2[84]
Scenedesmus sp. + Chlorella sp.:food waste1:4, C/N: 26.42Mesophilic639.832.8[99]
Nannochloropsis gaditana:cellulose1:3, C/N: 20.32372681.5[11]
Scenedesmus sp and C. vulgaris:sewage sludge12:88, C/N: 9.42373880[45]
Dunaliella salina:Olive mill solid waste1:3, C/N: 26.723733028.7[28]
Nannochloropsis salina:corn silage1:6, C/N: 21.22.74066015[60]

Table 4.

Methane yield of co-digested microalgae with different substrates at different ratios in BMP assays.

The above prerequisites have been selected as they are crucial for comparison purposes. The authors of this review acknowledge the lack of homogenization regarding BMP performance, which difficult the task to evaluate scientific data and provide reliable conclusions. The authors of this review also acknowledge the lack of a widely approved and used standard methodology for BMP tests, although, some are published and proved to be reliable enough [111, 112]. The authors of this review would also like to highlight that only 20 out of 120 reports complies with the above prerequisites, and if other crucial factors were included within the prerequisites (e.g. use of positive control or details on inoculum acclimation) no reports could have been included in this review. Additionally, when more than one co-substrate ratio was measured in the same study, only the one producing the highest methane yield was included in Table 4.

As can be observed in Table 4, methane yield ranged from 237.1 to 639.8NLCH4kgVS1. AcoD of microalgae with industrial wastes showed positive synergetic effects in most cases, being the improvement against the theoretical values up to 37%. However, some studies had reported negative effects. This could be due to the limits of batch methodologies and could lead to higher methane productions during continuous tests as pointed out by several authors [16, 33] or could be related to the low C/N ratio [45].

The optimum C/N ratio produced higher synergetic effects than those studies where the C/N ratio was above 30 or below 20. The most common and successful temperature is within the mesophilic range with an ISR of 2–3, although, some studies have reported improvements when applying thermophilic temperature at a higher ISR (4–5) [40, 77].

Regarding the co-substrates ratio, when microalgae were used as a nitrogen source in order to balance the C/N ratio, it was added commonly as a fourth part of the whole influent, or even less. When microalgae were co-digested with other low C/N substrates, it was added at higher concentrations. Nevertheless, the optimum ratio between microalgae and other co-substrates is unique and needs to be assessed through experimentation.

Advertisement

7. Scaling up the AcoD process

Table 5 summarizes all the results obtained by semi-continuous or continuous assays at lab or pilot scale results up to date complying with the following prerequisites:

  • The co-substrates ratio is well described as VS or C/N.

  • Temperature, OLR, and HRT/ISR has been considered and reported.

  • Methane yield have been reported under standard conditions (when values were reported under normal conditions a factor of 0.8871 was applied to obtain the methane yield under standard conditions).

Co-substrateCo-substrate ratio (VS)SystemAcclimated inoculumConditionsMethane yieldImprovement over controlRef.
Microalgae biomass:WW26.4:1 (VSS)UASB (pilot scale)Yes25 C
OLR: 1.0 g COD L−1 d−1
HRT: 8.1 h
235 NL kgVS−1−20% (wastewater)[32]
Chlorella sp.:WW38:62, C/N: 7.08AnMBR (lab scale)Yes25 C
OLR: 0.5 g COD L−1 d−1
HRT: 30 days
HRT: 70 days
391 NL kgVS−1nd*[62]
Scenedesmus sp. Chlorella sp.:WW38:62, C/N: 7.08AnMBR (pilot scale)Yes35°C
OLR: 0.5 g COD L−1 d−1
HRT: 30 days
HRT: 70 days
370 NL kgVS−1nd*[62]
Chlorella sp.:WW38:62, C/N: 7.08CSTR (lab scale)Yes55°C
OLR: 0.5 g COD L−1 d−1
HRT: 30 days
242 NL kg VS−1nd*[61]
Microalgae biomass:sewage(Flow rate) algae: 0.5 L h−1, sewage: 49 L h−1UASB (pilot scale)No23°C
OLR: 0.7 g VS L−1 d−1
HRT: 7 h
211 NL kgVS−134.8% (Sewage only)[79]
Chlorella sp.:WW38:62, C/N: 7.08Two stage: AnMBR, AnR (lab scale)NoAnMBR
35°C
OLR: 0.52 g COD L−1 d−1
HRT: 30 d,
AnR
35°C
OLR: 0.15 g COD L−1 d−1
HRT: 100 days
291 NL kgCOD−1nd*[65]
Scenedesmus sp.:primary sludge38:62AnMBR (lab scale)Yes35 C
OLR: 0.5 g COD L−1 d−1
HRT: 30 d
241 NL kgCOD−140.1% (algae only)[93]
Chlorella sp.:primary sludge38:62AnMBR (lab scale)Yes35 C
OLR: 0.5 g COD L−1 d−1
HRT: 30 d
228 NL kgCOD−16.5% (algae only)[93]
Chlorella vulgaris:chicken litter:glycerol30:67:3, C/N: 8.15Semi-continuous (lab scale)Yes37 C
OLR: 0.716 g VS L−1 d−1
HRT: 30 d
240 NL kgVS−139% (chicken litter only)[43]
Nannochloropsis limnetica:piggery slurry2:3CSTR (lab scale)nd*53 C
OLR: 1.4 g VS L−1 d−1
HRT: 15 d
216 NL kgVS−122.7% (pig slurry only);
−31% (algae only)
[77]
Microalgae biomass:WW1:3Semi-continuous (lab scale)No37 C
OLR: 1.17 g VS L−1 d−1
HRT: 30 days
460 NL kgVS−1187.5% (pretreated algae only)[74]
Stigeoclonium Scenedesmus:papaya waste1:1 (w/w)Semi-continuous (lab scale)Yes35 C
OLR: 1.1 g VS L−1 d−1
HRT: 31 days
230 NL kgCOD−159.7% (algae only); 12.2% (theoretical)[17]
Microalgae biomass:primary sludge:WAS2:3, C/N: 8.49Semi-continuous (lab scale)nd*37 C
OLR: 2.4 g COD L−1 d−1
HRT: 15 days
168 NL kgVS−1−15.8% (WWTP only)[46]
Chlorella sp.:wheat straw1:1, C/N: 13.1Semi-continuous (lab scale)nd*37 C
OLR: 1 g VS L−1 d−1
HRT: 20 days
240 NL kgVS−115% (no pretreated); 75% (algae only)[71]
Chlorella sp.:swine manure6:94Semi-continuous (lab scale)nd*35 C
OLR: 1.16–1.68 g VS L−1 d−1
HRT: 21 days
348 NL kgVS−19.8% (swine only)[84]
Chlorella vulgaris (MACC-755):used cooking oil1:1, C/N: 477Semi-continuous (lab scale)No38 C
OLR: 4.01 g VS L−1 d−1
HRT: 90 days
880 NL kgVS−1 2.86 L L−1 d−16.0% (algae only)[54]
Chlorella vulgaris (MACC-755):maize silage1:1, C/N: 16Semi-continuous (lab scale)No38 C
OLR: 4.01 g VS L−1 d−1
HRT: 90 days
1.99 L L−1 d−1nd*[54]
Chlorella vulgaris (MACC-755):mill residue1:1, C/N: 12Semi-continuous (lab scale)No38 C
OLR: 4.01 g VS L−1 d−1
HRT: 90 days
1.96 L L−1 d−1nd*[54]

Table 5.

Methane yield of co-digested microalgae with different substrates at different ratios in semi-continuous or continuous assays.

nd: not determined.


Semi-continuous or continuous processes could overpass some drawbacks from batch assays as reported by several authors [71, 77, 84]. This could be due to the acclimation of the inoculum during the experiments. To this sense, these processes had shown successful results in the co-digestion of substrates with low C/N ratios, low pH, and higher ammonium content, allowing a higher concentration of microalgae in the influent. However, some studies had shown negative results when compared with the AD of sole substrates [32, 46, 77]. This low methane production had been related to the difficulty of microalgae cell-wall digestion, accumulation over time of ammonium, or high OLR [32, 46, 77].

Regarding operational parameters, the most common used temperature was within the mesophilic range due to its low energy cost and good performance. OLR ranged from 0.5 to 4 g COD L−1 d−1, although, most studies showed a higher methane production when the OLR is below 1.5 g COD L−1 d−1. HRT was around 30 days, with some studies reporting 90 days when the OLR is around 4 g COD L−1 d−1. OLR and HRT had been discussed as key factors for a complete energy-effective harnessing of microalgae AcoD, to this sense, high OLR and low HRT would be the most effective way for achieving this. However, pH decreases due to VFA accumulation, and ammonium release are the main factors affecting the OLR.

Advertisement

8. Conclusions

The status and current trends of AcoD of microalgae and industrial wastes were reviewed in this chapter. AcoD performance improvements still need further research on varied co-substrates and optimal mix ratios. Operational parameters and their control are key to achieving optimal biogas. Pretreatments of microalgae biomass are a promising way to enhance biogas production. The majority of research investigations are done by a few research groups and centered on biomethane potential tests lacking a common methodology, thus, further research and the application of common criteria need to be implemented. Moreover, pilot-scale assays have shown promising results, however, very few research groups have the ability to implement these studies.

Advertisement

Acknowledgments

The authors wish to express their gratitude to the regional government of Andalucía, Junta de Andalucía, Consejería de Transformación Económica, Industria, Conocimiento y Universidades (Project FEDER UPO-1380782) and to the Spanish national R&D plan (PID2020-114975RB-100/AEI/10.13039/5011000/11033) for providing financial support.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Vargas-Estrada L, Longoria A, Arenas E, Moreira J, Okoye PU, Bustos-Terrones Y, et al. A review on current trends in biogas production from miroalgae biomass and microalgae waste by anaerobic digestion and co-digestion. Bioenergy Research. 2021;15:77-92
  2. 2. Thorin E, Olsson J, Schwede S, Nehrenheim E. Biogas from co-digestion of sewage sludge and microalgae. Energy Procedia. 2017;105:1037-1042
  3. 3. Ganesh Saratale R, Kumar G, Banu R, Xia A, Periyasamy S, Dattatraya SG. A critical review on anaerobic digestion of microalgae and macroalgae and co-digestion of biomass for enhanced methane generation. Bioresource Technology. 2018;262:319-332
  4. 4. Choudhary P, Assemany PP, Naaz F, Bhattacharya A, Castro J, Couto E, et al. A review of biochemical and thermochemical energy conversion routes of wastewater grown algal biomass. Science of the Total Environment. 2020;726:137961
  5. 5. Wall DM, McDonagh S, Murphy JD. Cascading biomethane energy systems for sustainable green gas production in a circular economy. Bioresource Technology. 2017;243:1207-1215
  6. 6. Thorin E, Olsson J, Schwede S, Nehrenheim E. Co-digestion of sewage sludge and microalgae—Biogas production investigations. Applied Energy. 2018;227:64-72
  7. 7. Calicioglu O, Demirer GN, Ganesh Saratale R, Kumar G, Banu R, Xia A, et al. Recent achievements in the production of biogas from microalgae. Waste and Biomass Valorization. 2017;8:129-139
  8. 8. Agustini CB, da Fontoura JT, Mella B, Gutterres M. Evaluating co-substrates to supplement biogas production from tannery solid waste treatment—Cattle hair, microalgae biomass, and silicone. Biofuels, Bioproducts and Biorefining. 2018;12:1095-1102
  9. 9. Ahmad A, Shah SMU, Buang A, Othman MF, Abdullah MA. Evaluation of aerobic and anaerobic co-digestion of Tetraselmis suecica and oil palm empty fruit bunches by response surface methodology. Advanced Materials Research. 2014;925:243-247
  10. 10. Avila R, Justo Á, Carrero E, Crivillés E, Vicent T, Blánquez P. Water resource recovery coupling microalgae wastewater treatment and sludge co-digestion for bio-wastes valorisation at industrial pilot-scale. Bioresource Technology. 2022;343:126080
  11. 11. Barontini F, Biagini E, Dragoni F, Corneli E, Ragaglini G, Bonari E, et al. Anaerobic digestion and co-digestion of oleaginous microalgae residues for biogas production. Chemical Engineering Transactions. 2016;50:91-96
  12. 12. Beltrán C, Jeison D, Fermoso FG, Borja R. Batch anaerobic co-digestion of waste activated sludge and microalgae (Chlorella sorokiniana) at mesophilic temperature. Journal of Environmental Science and Health—Part A Toxic/Hazardous Substances and Environmental Engineering. 2016;51:847-850
  13. 13. Calicioglu O, Demirer GN. Carbon-to-nitrogen and substrate-to-inoculum ratio adjustments can improve co-digestion performance of microalgal biomass obtained from domestic wastewater treatment. Environmental Technology (United Kingdom). 2019;40:614-624
  14. 14. Caporgno MP, Clavero E, Torras C, Salvadó J, Lepine O, Pruvost J, et al. Energy and nutrients recovery from lipid-extracted nannochloropsis via anaerobic digestion and hydrothermal liquefaction. ACS Sustainable Chemistry and Engineering. 2016;4:3133-3139
  15. 15. Caporgno MP, Olkiewicz M, Torras C, Salvadó J, Clavero E, Bengoa C. Effect of pre-treatments on the production of biofuels from Phaeodactylum tricornutum. Journal of Environmental Management. 2016;177:240-246
  16. 16. Carminati P, Gusmini D, Pizzera A, Catenacci A, Parati K, Ficara E. Biogas from mono- and co-digestion of microalgal biomass grown on piggery wastewater. Water Science and Technology. 2018;78:103-113
  17. 17. Cea-Barcia G, Pérez J, Buitrón G. Co-digestion of microalga-bacteria biomass with papaya waste for methane production. Water Science and Technology. 2018;78:125-131
  18. 18. Cheng Q , Deng F, Li H, Qin Z, Wang M, Li J. Nutrients removal from the secondary effluents of municipal domestic wastewater by Oscillatoria tenuis and subsequent co-digestion with pig manure. Environmental Technology (United Kingdom). 2018;39:3127-3134
  19. 19. Damtie MM, Shin J, Jang HM, Kim YM. Synergistic co-digestion of microalgae and primary sludge to enhance methane yield from temperature-phased anaerobic digestion. Energies. 2020;13:4547
  20. 20. Ahmad A, Shah SMU, Othman MF, Abdullah MA. Enhanced palm oil mill effluent treatment and biomethane production by co-digestion of oil palm empty fruit bunches with Chlorella sp. Canadian Journal of Chemical Engineering. 2014;92:1636-1642
  21. 21. Dębowski M, Zieliński M, Kisielewska M, Krzemieniewski M. Anaerobic co-digestion of the energy crop Sida hermaphrodita and microalgae biomass for enhanced biogas production. International Journal of Environmental Research. 2017;11:243-250
  22. 22. Du X, Tao Y, Li H, Liu Y, Feng K. Synergistic methane production from the anaerobic co-digestion of Spirulina platensis with food waste and sewage sludge at high solid concentrations. Renewable Energy. 2019;142:55-61
  23. 23. Du X, Tao Y, Liu Y, Li H. Stimulating methane production from microalgae by alkaline pretreatment and co-digestion with sludge. Environmental Technology (United Kingdom). 2020;41:1546-1553
  24. 24. Ehimen EA, Sun ZF, Carrington CG, Birch EJ, Eaton-Rye JJ. Anaerobic digestion of microalgae residues resulting from the biodiesel production process. Applied Energy. 2011;88:3454-3463
  25. 25. Fernández-Rodríguez MJ, de la Lama-Calvente D, Jiménez-Rodríguez A, Borja R, Rincón B. Evolution of control parameters in biochemical methane potential tests of olive mill solid waste (OMSW), thermal pre-treated OMSW, and its co-digestion with Dunaliella salina. Journal of Applied Phycology. 2021;33:419-429
  26. 26. Fernández-Rodríguez MJ, de la Lama-Calvente D, Jiménez-Rodríguez A, Borja R, Rincón-Llorente B. Anaerobic co-digestion of olive mill solid waste and microalga Scenedesmus quadricauda: Effect of different carbon to nitrogen ratios on process performance and kinetics. Journal of Applied Phycology. 2019;31:3583-3591
  27. 27. Fernández-Rodríguez MJ, de la Lama-Calvente D, Jiménez-Rodríguez A, Borja R, Rincón-Llorente B. Influence of the cell wall of Chlamydomonas reinhardtii on anaerobic digestion yield and on its anaerobic co-digestion with a carbon-rich substrate. Process Safety and Environmental Protection. 2019;128:167-175
  28. 28. Fernández-Rodríguez MJ, Rincón B, Fermoso FG, Jiménez AM, Borja R. Assessment of two-phase olive mill solid waste and microalgae co-digestion to improve methane production and process kinetics. Bioresource Technology. 2014;157:263-269
  29. 29. Ferreira LO, Astals S, Passos F. Anaerobic co-digestion of food waste and microalgae in an integrated treatment plant. Journal of Chemical Technology and Biotechnology. 2021
  30. 30. Garoma T, Nguyen D. Anaerobic co-digestion of microalgae Scenedesmus sp. and TWAS for biomethane production. Water Environment Research. 2016;88:13-20
  31. 31. Ahmed D, Wagdy R, Said N. Evaluation of biogas production from anaerobic co-digestion of sewage sludge with microalgae and agriculture wastes. BioResources. 2019;14:8405-8412
  32. 32. Gonçalves RF, Assis TI, Maciel GB, Borges RM, Cassini STA. Co-digestion of municipal wastewater and microalgae biomass in an upflow anaerobic sludge blanket reactor. Algal Research. 2020;52:102117
  33. 33. Herrmann C, Kalita N, Wall D, Xia A, Murphy JD, Rétfalvi T, et al. Optimised biogas production from microalgae through co-digestion with carbon-rich co-substrates. Bioresource Technology. 2016;28:2741-2752
  34. 34. Hu Y, Kumar M, Wang Z, Zhan X, Stengel DB. Filamentous microalgae as an advantageous co-substrate for enhanced methane production and digestate dewaterability in anaerobic co-digestion of pig manure. Waste Management. 2021;119:399-407
  35. 35. Kim J, Kang C-M. Increased anaerobic production of methane by co-digestion of sludge with microalgal biomass and food waste leachate. Bioresource Technology. 2015;189:409-412
  36. 36. Kumar P, Bhattacharya A, Prajapati SK, Malik A, Vijay VK. Anaerobic co-digestion of waste microalgal biomass with cattle dung in a pilot-scale reactor: Effect of seasonal variations and long-term stability assessment. Biomass Conversion and Biorefinery. 2020;12:1203-1215
  37. 37. Li R, Duan N, Zhang Y, Liu Z, Li B, Zhang D, et al. Anaerobic co-digestion of chicken manure and microalgae Chlorella sp.: Methane potential, microbial diversity and synergistic impact evaluation. Waste Management. 2017;68:120-127
  38. 38. Li R, Duan N, Zhang Y, Liu Z, Li B, Zhang D, et al. Co-digestion of chicken manure and microalgae Chlorella 1067 grown in the recycled digestate: Nutrients reuse and biogas enhancement. Waste Management. 2017;70:247-254
  39. 39. Lu D, Liu X, Apul OG, Zhang L, Ryan DK, Zhang X. Optimization of biomethane production from anaerobic co-digestion of microalgae and septic tank sludge. Biomass and Bioenergy. 2019;127:105266
  40. 40. Mahdy A, Fotidis IA, Mancini E, Ballesteros M, González-Fernández C, Angelidaki I. Ammonia tolerant inocula provide a good base for anaerobic digestion of microalgae in third generation biogas process. Bioresource Technology. 2017;225:272-278
  41. 41. Meneses-Reyes JC, Hernández-Eugenio G, Huber DH, Balagurusamy N, Espinosa-Solares T. Biochemical methane potential of oil-extracted microalgae and glycerol in co-digestion with chicken litter. Bioresource Technology. 2017;224:373-379
  42. 42. Álvarez X, Arévalo O, Salvador M, Mercado I, Velázquez-Martí B. Cyanobacterial biomass produced in the wastewater of the dairy industry and its evaluation in anaerobic co-digestion with cattle manure for enhanced methane production. PRO. 2020;8:1-16
  43. 43. Meneses-Reyes JC, Hernández-Eugenio G, Huber DH, Balagurusamy N, Espinosa-Solares T. Oil-extracted Chlorella vulgaris biomass and glycerol bioconversion to methane via continuous anaerobic co-digestion with chicken litter. Renewable Energy. 2018;128:223-229
  44. 44. Neumann P, Torres A, Fermoso FG, Borja R, Jeison D. Anaerobic co-digestion of lipid-spent microalgae with waste activated sludge and glycerol in batch mode. International Biodeterioration and Biodegradation. 2015;100:85-88
  45. 45. Olsson J, Feng XM, Ascue J, Gentili FG, Shabiimam MA, Nehrenheim E, et al. Co-digestion of cultivated microalgae and sewage sludge from municipal waste water treatment. Bioresource Technology. 2014;171:203-210. DOI: 10.1016/j.biortech.2014.08.069
  46. 46. Olsson J, Forkman T, Gentili FG, Zambrano J, Schwede S, Thorin E, et al. Anaerobic co-digestion of sludge and microalgae grown in municipal wastewater—A feasibility study. Water Science and Technology. 2018;77:682-694
  47. 47. Panyaping K, Khiewwijit R, Wongpankamol P. Enhanced biogas production potential of microalgae and swine wastewater using co-digestion and alkaline pretreatment. Water Science and Technology. 2018;78:92-102
  48. 48. Park S, Li Y. Evaluation of methane production and macronutrient degradation in the anaerobic co-digestion of algae biomass residue and lipid waste. Bioresource Technology. 2012;111:42-48
  49. 49. Passos F, Cordeiro PHM, Baeta BEL, de Aquino SF, Perez-Elvira SI. Anaerobic co-digestion of coffee husks and microalgal biomass after thermal hydrolysis. Bioresource Technology. 2018;253:49-54
  50. 50. Peng S, Colosi LM. Anaerobic digestion of algae biomass to produce energy during wastewater treatment. Water Environment Research. 2016;88:29-39
  51. 51. Quintana-Najera J, Blacker AJ, Fletcher LA, Ross AB. Influence of augmentation of biochar during anaerobic co-digestion of Chlorella vulgaris and cellulose. Bioresource Technology. 2022;343:126086
  52. 52. Ramos-Suárez JL, Carreras N. Use of microalgae residues for biogas production. Chemical Engineering Journal. 2014;242:86-95
  53. 53. Ambarsari H, Adrian R, Manurung BS. Anaerobic biogas production using microalgae Chlorella sp. as biomass co-digested by cow manure and cow rumen fluid as inoculum. IOP Conference Series: Earth and Environmental Science. 2018
  54. 54. Rétfalvi T, Szabó P, Hájos AT, Albert L, Kovács A, Milics G, et al. Effect of co-substrate feeding on methane yield of anaerobic digestion of Chlorella vulgaris. Journal of Applied Phycology. 2016;28:2741-2752
  55. 55. Rincón-Pérez J, Celis LB, Morales M, Alatriste-Mondragón F, Tapia-Rodríguez A, Razo-Flores E. Improvement of methane production at alkaline and neutral pH from anaerobic co-digestion of microalgal biomass and cheese whey. Biochemical Engineering Journal. 2021;169:107972
  56. 56. Rusten B, Sahu AK. Microalgae growth for nutrient recovery from sludge liquor and production of renewable bioenergy. Water Science and Technology. 2011;64:1195-1201
  57. 57. Saleem M, Hanif MU, Bahada A, Iqbal H, Capareda SC, Waqas A. The effects of hotwater and ultrasonication pretreatment of microalgae (Nannochloropsis oculata) on biogas production in anaerobic co-digestion with cow manure. PRO. 2020;8:1-10
  58. 58. Santos-Ballardo DU, Font-Segura X, Ferrer AS, Barrena R, Rossi S, Valdez-Ortiz A. Valorisation of biodiesel production wastes: Anaerobic digestion of residual Tetraselmis suecica biomass and co-digestion with glycerol. Waste Management and Research. 2015;33:250-257
  59. 59. Scarcelli PG, Serejo ML, Paulo PL, Boncz MÁ. Evaluation of biomethanization during co-digestion of thermally pretreated microalgae and waste activated sludge, and estimation of its kinetic parameters. Science of the Total Environment. 2020;706:135745
  60. 60. Schwede S, Kowalczyk A, Gerber M, Span R. Anaerobic co-digestion of the marine microalga Nannochloropsis salina with energy crops. Bioresource Technology. 2013;148:428-435. DOI: 10.1016/j.biortech.2013.08.157
  61. 61. Serna-García R, Borrás L, Bouzas A, Seco A. Insights into the biological process performance and microbial diversity during thermophilic microalgae co-digestion in an anaerobic membrane bioreactor (AnMBR). Algal Research. 2020;50:101981
  62. 62. Serna-García R, Mora-Sánchez JF, Sanchis-Perucho P, Bouzas A, Seco A. Anaerobic membrane bioreactor (AnMBR) scale-up from laboratory to pilot-scale for microalgae and primary sludge co-digestion: Biological and filtration assessment. Bioresource Technology. 2020;316:123930
  63. 63. Serna-García R, Ruiz-Barriga P, Noriega-Hevia G, Serralta J, Pachés M, Bouzas A. Maximising resource recovery from wastewater grown microalgae and primary sludge in an anaerobic membrane co-digestion pilot plant coupled to a composting process. Journal of Environmental Management. 2021;281:111890
  64. 64. Arias DM, Solé-Bundó M, Garfí M, Ferrer I, García J, Uggetti E. Integrating microalgae tertiary treatment into activated sludge systems for energy and nutrients recovery from wastewater. Bioresource Technology. 2018;247:513-519
  65. 65. Serna-García R, Zamorano-López N, Seco A, Bouzas A. Co-digestion of harvested microalgae and primary sludge in a mesophilic anaerobic membrane bioreactor (AnMBR): Methane potential and microbial diversity. Bioresource Technology. 2020;298
  66. 66. Sialve B, Bernet N, Bernard O. Anaerobic digestion of microalgae as a necessary step to make microalgal biodiesel sustainable. Biotechnology Advances. 2009;27:409-416
  67. 67. Siddique MNI, Wahid ZBA. Enhanced methane yield by codigestion of sewage sludge with microalgae and catering waste leachate. Water Environment Research. 2018;90:835-839
  68. 68. Sitthikitpanya N, Sittijunda S, Khamtib S, Reungsang A. Co-generation of biohydrogen and biochemicals from co-digestion of Chlorella sp. biomass hydrolysate with sugarcane leaf hydrolysate in an integrated circular biorefinery concept. Biotechnology for Biofuels. 2021;14:197
  69. 69. Sittijunda S, Reungsang A. Methane production from the co-digestion of algal biomass with crude glycerol by anaerobic mixed cultures. Waste and Biomass Valorization. 2020;11:1873-1881
  70. 70. Solé-Bundó M, Cucina M, Folch M, Tàpias J, Gigliotti G, Garfí M, et al. Assessing the agricultural reuse of the digestate from microalgae anaerobic digestion and co-digestion with sewage sludge. Science of the Total Environment. 2017;586:1-9
  71. 71. Solé-Bundó M, Eskicioglu C, Garfí M, Carrère H, Ferrer I. Anaerobic co-digestion of microalgal biomass and wheat straw with and without thermo-alkaline pretreatment. Bioresource Technology. 2017;237:89-98
  72. 72. Solé-Bundó M, Garfí M, Ferrer I. Pretreatment and co-digestion of microalgae, sludge and fat oil and grease (FOG) from microalgae-based wastewater treatment plants. Bioresource Technology. 2020;298:122563
  73. 73. Solé-Bundó M, Garfí M, Matamoros V, Ferrer I. Co-digestion of microalgae and primary sludge: Effect on biogas production and microcontaminants removal. Science of the Total Environment. 2019;660:974-981
  74. 74. Solé-Bundó M, Salvadó H, Passos F, Garfí M, Ferrer I. Strategies to optimize microalgae conversion to biogas: Co-digestion, pretreatment and hydraulic retention time. Molecules. 2018;23:2096
  75. 75. Astals S, Musenze RS, Bai X, Tannock S, Tait S, Pratt S, et al. Anaerobic co-digestion of pig manure and algae: Impact of intracellular algal products recovery on co-digestion performance. Bioresource Technology. 2015;181:97-104
  76. 76. Srivastava G, Kumar V, Tiwari R, Patil R, Kalamdhad A, Goud V. Anaerobic co-digestion of defatted microalgae residue and rice straw as an emerging trend for waste utilization and sustainable biorefinery development. Biomass Conversion and Biorefinery. 2020;12:1193-1202
  77. 77. Tsapekos P, Kougias PG, Alvarado-Morales M, Kovalovszki A, Corbière M, Angelidaki I. Energy recovery from wastewater microalgae through anaerobic digestion process: Methane potential, continuous reactor operation and modelling aspects. Biochemical Engineering Journal. 2018;139:1-7
  78. 78. Varol A, Ugurlu A. Biogas production from microalgae (Spirulina platensis) in a two stage anaerobic system. Waste and Biomass Valorization. 2016;7:193-200
  79. 79. Vassalle L, Díez-Montero R, Machado ATR, Moreira C, Ferrer I, Mota CR, et al. Upflow anaerobic sludge blanket in microalgae-based sewage treatment: Co-digestion for improving biogas production. Bioresource Technology. 2020;300:122677
  80. 80. Vassalle L, Passos F, Rosa-Machado AT, Moreira C, Reis M, Pascoal de Freitas M, et al. The use of solar pre-treatment as a strategy to improve the anaerobic biodegradability of microalgal biomass in co-digestion with sewage. Chemosphere. 2022;286:131929
  81. 81. Wágner DS, Radovici M, Smets BF, Angelidaki I, Valverde-Pérez B, Plósz BG. Harvesting microalgae using activated sludge can decrease polymer dosing and enhance methane production via co-digestion in a bacterial-microalgal process. Algal Research. 2016;20:197-204
  82. 82. Wang M, Kuo-Dahab WC, Park C. Investigation of characteristics of microalgae grown in different wastewater and their enhancing anaerobic digestibility of waste activated sludge. 86th Annual Water Environment Federation Technical Exhibition and Conference, WEFTEC 2013. 2013. p. 2071-2080
  83. 83. Wang M, Lee E, Zhang Q , Ergas S. Energy production from anaerobic co-digestion of swine manure and microalgae Chlorella sp. 87th Annual Water Environment Federation Technical Exhibition and Conference, WEFTEC 2014. 2014. p. 3806-3816
  84. 84. Wang M, Lee E, Zhang Q , Ergas SJ. Anaerobic co-digestion of swine manure and microalgae Chlorella sp.: Experimental studies and energy analysis. Bioenergy Research. 2016;9:1204-1215
  85. 85. Wang M, Park C. Investigation of anaerobic digestion of Chlorella sp. and Micractinium sp. grown in high-nitrogen wastewater and their co-digestion with waste activated sludge. Biomass and Bioenergy. 2015;80:30-37
  86. 86. Avila R, Carrero E, Crivillés E, Mercader M, Vicent T, Blánquez P. Effects of low temperature thermal pretreatments in solubility and co-digestion of waste activated sludge and microalgae mixtures. Algal Research. 2020;50:101965
  87. 87. Wang M, Sahu AK, Rusten B, Park C. Anaerobic co-digestion of microalgae Chlorella sp. and waste activated sludge. Bioresource Technology. 2013;142:585-590
  88. 88. Wang M, Zhu Z, Dolan S, Park C. Investigation of algal cultivation and anaerobic co-digestion of sewage sludge and algae at wastewater treatment plant (WWTP). WEFTEC 2012—85th Annual Technical Exhibition and Conference. 2012. p. 3586-3599
  89. 89. Wirth R, Lakatos G, Böjti T, Maróti G, Bagi Z, Kis M, et al. Metagenome changes in the mesophilic biogas-producing community during fermentation of the green alga Scenedesmus obliquus. Journal of Biotechnology. 2015;215:52-61
  90. 90. Wirth R, Pap B, Dudits D, Kakuk B, Bagi Z, Shetty P, et al. Genome-centric investigation of anaerobic digestion using sustainable second and third generation substrates. Journal of Biotechnology. 2021;339:53-64
  91. 91. Woinaroschy A, Matei CB, Gǎlan A-M, Stepan E. Anaerobic co-digestion of delipidized microalgae biomass and food residues. UPB Scientific Bulletin, Series B: Chemistry and Materials Science. 2018;80:41-52
  92. 92. Yuan X, Wang M, Park C, Sahu AK, Ergas SJ. Microalgae growth using high-strength wastewater followed by anaerobic co-digestion. Water Environment Research. 2012;84:396-404
  93. 93. Zamorano-López N, Borrás L, Seco A, Aguado D. Unveiling microbial structures during raw microalgae digestion and co-digestion with primary sludge to produce biogas using semi-continuous AnMBR systems. Science of the Total Environment. 2020;699:134365
  94. 94. Zhang Y, Caldwell GS, Blythe PT, Zealand AM, Li S, Edwards S, et al. Co-digestion of microalgae with potato processing waste and glycerol: Effect of glycerol addition on methane production and the microbial community. RSC Advances. 2020;10:37391-37408
  95. 95. Zhang Y, Caldwell GS, Sallis PJ. Semi-continuous anaerobic co-digestion of marine microalgae with potato processing waste for methane production. Journal of Environmental Chemical Engineering. 2019;7:102917
  96. 96. Zhang Y, Caldwell GS, Zealand AM, Sallis PJ. Anaerobic co-digestion of microalgae Chlorella vulgaris and potato processing waste: Effect of mixing ratio, waste type and substrate to inoculum ratio. Biochemical Engineering Journal. 2019;143:91-100
  97. 97. Avila R, Carrero E, Vicent T, Blánquez P. Integration of enzymatic pretreatment and sludge co-digestion in biogas production from microalgae. Waste Management. 2021;124:254-263
  98. 98. Zhang Y, Kang X, Wang Z, Kong X, Li L, Sun Y, et al. Enhancement of the energy yield from microalgae via enzymatic pretreatment and anaerobic co-digestion. Energy. 2018;164:400-407
  99. 99. Zhen G, Lu X, Kobayashi T, Kumar G, Xu K. Anaerobic co-digestion on improving methane production from mixed microalgae (Scenedesmus sp., Chlorella sp.) and food waste: Kinetic modeling and synergistic impact evaluation. Chemical Engineering Journal. 2016;299:332-341
  100. 100. Wannapokin A, Ramaraj R, Whangchai K, Unpaprom Y. Potential improvement of biogas production from fallen teak leaves with co-digestion of microalgae. 3 Biotech. 2018;8:123
  101. 101. Golueke CG, Oswald WJ, Gotaas HB. Anaerobic Digestion of Algae. Applied Microbiology. 1957;5(1):47-55
  102. 102. Zabed HM, Akter S, Yun J, Zhang G, Zhang Y, Qi X. Biogas from microalgae: Technologies, challenges and opportunities. Renewable and Sustainable Energy Reviews. 2020;117:109503
  103. 103. Samson R, Leduy A. Improved performance of anaerobic digestion of Spirulina maxima algal biomass by addition of carbon-rich subtrates. Biotechnology Letters. 1983;5(10):677-682
  104. 104. Selormey GK, Barnes B, Kemausuor F, Darkwah L. A review of anaerobic digestion of slaughterhouse waste: effect of selected operational and environmental parameters on anaerobic biodegradability. Reviews in Environmental Science and Biotechnology. 2021;20:1073-1086
  105. 105. Kabouris JC, Tezel U, Pavlostathis SG, Engelmann M, Dulaney JA, Todd AC, et al. Mesophilic and thermophilic anaerobic digestion of municipal sludge and fat, oil, and grease. Water Environment Research. 2009;81:476-485
  106. 106. Younus Bhuiyan Sabbir ASM, Saha CK, Nandi R, Zaman MFU, Alam MM, Sarker S. Effects of seasonal temperature variation on slurry temperature and biogas composition of a commercial fixed-dome anaerobic digester used in Bangladesh. Sustainability (Switzerland). 2021;13:11096
  107. 107. Zhang T, Mao C, Zhai N, Wang X, Yang G. Influence of initial pH on thermophilic anaerobic co-digestion of swine manure and maize stalk. Waste Management. 2015;35:119-126
  108. 108. Cai Y, Gallegos D, Zheng Z, Stinner W, Wang X, Pröter J, et al. Exploring the combined effect of total ammonia nitrogen, pH and temperature on anaerobic digestion of chicken manure using response surface methodology and two kinetic models. Bioresource Technology. 2021;337:125328
  109. 109. Carotenuto C, Guarino G, Morrone B, Minale M. Temperature and pH effect on methane production from buffalo manure anaerobic digestion. International Journal of Heat and Technology. 2016;34:S425-S429
  110. 110. Ramos-Suárez JL, Martínez A, Carreras N. Optimization of the digestion process of Scenedesmus sp. and Opuntia maxima for biogas production. Energy Conversion and Management. 2014;88:1263-1270
  111. 111. Raposo F, Borja R, Ibelli-Bianco C. Predictive regression models for biochemical methane potential tests of biomass samples: Pitfalls and challenges of laboratory measurements. Renewable and Sustainable Energy Reviews. 2020;127:109890
  112. 112. Holliger C, Alves M, Andrade D, Angelidaki I, Astals S, Baier U, et al. Towards a standardization of biomethane potential tests. Water Science and Technology. 2016;74:2515-2522

Written By

David de la Lama-Calvente, Juan Cubero, María José Fernández-Rodríguez, Antonia Jiménez-Rodríguez and Rafael Borja

Submitted: 21 February 2022 Reviewed: 07 March 2022 Published: 19 May 2022