Wang2016 Microalgal For Insdustrial Wastewater Treatment

Download as pdf or txt
Download as pdf or txt
You are on page 1of 52

Accepted Manuscript

Review

Perspectives on the feasibility of using microalgae for industrial wastewater


treatment

Yue Wang, Shih-Hsin Ho, Chieh-Lun Cheng, Wan-Qian Guo, Dillirani


Nagarajan, Nan-Qi Ren, Duu-Jong Lee, Jo-Shu Chang

PII: S0960-8524(16)31373-6
DOI: http://dx.doi.org/10.1016/j.biortech.2016.09.106
Reference: BITE 17129

To appear in: Bioresource Technology

Received Date: 4 September 2016


Revised Date: 25 September 2016
Accepted Date: 26 September 2016

Please cite this article as: Wang, Y., Ho, S-H., Cheng, C-L., Guo, W-Q., Nagarajan, D., Ren, N-Q., Lee, D-J., Chang,
J-S., Perspectives on the feasibility of using microalgae for industrial wastewater treatment, Bioresource
Technology (2016), doi: http://dx.doi.org/10.1016/j.biortech.2016.09.106

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
A revised manuscript (BITE-D-16-05062R1) submitted to Bioresource Technology

Perspectives on the feasibility of using microalgae for industrial


wastewater treatment

Yue Wang1, Shih-Hsin Ho 1, Chieh-Lun Cheng2, Wan-Qian Guo1, Dillirani Nagarajan2,


Nan-Qi Ren1, Duu-Jong Lee1,3, Jo-Shu Chang1,2,4*

1
State Key Laboratory of Urban Water Resource and Environment, Harbin Institute of
Technology, Harbin, China
2
Department of Chemical Engineering, National Cheng Kung University, Tainan,
Taiwan
3
Department of Chemical Engineering, National Taiwan University, Taipei, Taiwan
4
Research Center for Energy Technology and Strategy, National Cheng Kung
University, Tainan 701, Taiwan

*
Corresponding author: Professor Jo-Shu Chang; e-mail: [email protected]

BITE-D-16-05062R1 (revised manuscript).docx 10/12/2016

1
Abstract

Although microalgae can serve as an appropriate alternative feedstock for biofuel

production, the high microalgal cultivation cost has been a major obstacle for

commercializing such attempts. One of the feasible solution for cost reduction is to

couple microalgal biofuel production system with wastewater treatment, as

microalgae are known to effectively eliminate a variety of nutrients/pollutants in

wastewater, such as nitrogen/phosphate, organic carbons, VFAs, pharmaceutical

compounds, textile dye compounds, and heavy metals. This review aims to critically

discuss the feasibility of microalgae-based wastewater treatment, including the

strategies for strain selection, the effect of wastewater types, photobioreactor design,

economic feasibility assessment, and other key issues that influence the treatment

performance. The potential of microalgae-bacteria consortium for treatment of

industrial wastewaters is also discussed. This review provides useful information for

developing an integrated wastewater treatment with microalgal biomass and biofuel

production facilities and establishing efficient co-cultivation for microalgae and

bacteria in such systems.

Key words: microalgae, algae-bacteria consortium, wastewater treatment,

photobioreactor

2
1. Introduction

Global carbon emissions are expected to increase from the current 404 ppm of

atmospheric CO2, along with severe environmental issues. Algae are an effective

intermediary which can convert carbon dioxide and solar energy into various

bio-energy forms (such as biodiesel, bio-ethanol, and bio-butanol), since they possess

20% higher photosynthetic efficiency compared to terrestrial plants. However,

currently, the worldwide production of microalgal biomass is only about 9000 tons y-1 .

More importantly, the production cost is as high as $20-$200 kg-1, which significantly

hinders the use of microalgae as biofuel producer. Some scientists demonstrated that

algae cultivation using various wastewaters is a more effective and economic way to

reduce the biomass production cost, and notably, providing the extra benefits of

simultaneous wastewater treatment (Wang & Lan, 2011; Wang et al., 2015b; Wilkie &

Mulbry, 2002). It has been reported that some microalgae can simultaneously remove

nitrogen and phosphorus from domestic wastewater down to very low concentrations

of 2.2 mg L-1 and 0.15 mg L-1, respectively, by uptake of the nutrients into the cells

(Boelee et al., 2011). Until now, microalgae is most commonly used for the treatment

of municipal wastewater and concomitant biomass production. Some high-COD and

nutrient-rich wastewaters, such as industrial and agriculture wastewaters are usually

combined with anaerobic pretreatment or diluted appropriately to avoid algal growth

inhibition caused by the high COD concentrations (Wang et al., 2015b). In recent days,

an increased number of studies on industry wastewater treatment by microalgae are

also successfully conducted, such as sugar mill effluent , pulp and paper industry

3
effluent (Polishchuk et al., 2015), fish farm wastewater, coal-fired metal-contaminated

wastewater , petroleum industrial wastewater , pharmaceutical industry wastewater ,

textile dye industry effluent, and electroplating industry wastewater. Notably, both

photobioreactors and wastewater treatment plants (such as oxidation ponds) can be

employed for microalgal biomass production and simultaneous wastewater treatment.

High rate algal ponds (HRAPs) and oxidation ponds are the most employed

wastewater treatment plants for both municipal and industrial wastewater treatment

combined with microalgal biomass production. Considering the above mentioned

factors, the high potential of microalgal wastewater treatment process is clearly

demonstrated (Fig. 1).

However, few studies reported that the algae organic matters (AOM) secreted

during algal growth would form various odor and taste compounds, which will

adversely affect the effluent quality after algal growth (Wang et al., 2015b). Odor is an

index to evaluate the quality of effluent from wastewater treatment plant, and these

compounds can increase the COD of effluent. This is disadvantageous, as the primary

aim of wastewater treatment is to reduce COD and other organic nutritional

compounds in the effluent. Compared to the conventional activated sludge based

wastewater treatment, several obstacles in microalgae-based wastewater treatment

systems often exist: (1) The COD loading (F/M) is lower for algal treatment processes,

whereas activated sludge process can treat wastewaters with a wider COD range. For

instance, algae can grow well only in wastewater containing lower COD (below 5000

mg L-1) (He et al., 2013; Wang et al., 2015b). (2) The extracellular organic matters

4
produced during algal cultivation may increase the COD of wastewater (Wang et al.,

2015b). In general, the COD of wastewater will be consumed effectively in the first

several days, but in batch mode processes, the excess extracellular organic matters

may accumulate with prolonged cultivation time, resulting in a possible increase in

the COD level (Wang et al., 2015b). Thus, part of the microalgal biomass should be

harvested in time, to avoid the production of excess extracellular organic matters.

Most microalgae require longer period to accumulate huge amounts of lipid or

carbohydrate (i.e., used for biofuel production) and the risk of pollution and

ecological hazards caused by the AOM will significantly increase. (3) When

wastewater sources are located in areas with extreme environmental conditions, either

high altitude hypothermic area with low insolation (Talbot et al., 1991) or low altitude

high temperature region with high insolation , selection of an algae that can proliferate

under harsh environmental conditions is of paramount importance. However, most

microalgae species can only grow well within the range of 25 to 35 oC, which will

limit the use of microalgal wastewater treatment in practice (Ruiz-Martinez et al.,

2015).

Taking these into account, development of algae-bacteria consortium system may

be an ideal process to solve the above-mentioned obstacles associated with

microalgae-based wastewater treatment process. The AOM can help promote bacteria

zooglea formation in some cases since the AOM mainly appeared as polysaccharides.

In the co-cultivation system, the conventional activated sludge with bacteria can

effectively remove the organic carbon source (COD), along with the generation of

5
CO2. Through photosynthesis, algae can convert CO2 to biomass and produce O2 to

support the bacterial growth (Fig. 1). Thus, the algae-bacteria consortium system can

be perfectly applied for wastewater treatment to avoid the external oxygen supply,

allow nutrients assimilation into biomass, and reduce CO2 emissions to the

atmosphere. Particles of senescent and dead algal materials are sources of organic

compounds that are liberated as Dissolved Organic Matter (DOM) by aggregate

associated bacteria. Nevertheless, the implementation of an integrated algal-bacteria

system faces challenges as well. For examples: (1) Little is known about the complex

tripartite interactions between algae, their associated microbial consortia and

wastewater associated microbes for the growth and utilization stages of the biofuel

algae, leading to difficulties in creating a stable co-cultivation system. (2)

Phytoplankton-bacteria interactions can range from symbiotic to parasitic, with algae

inhibiting or benefitting bacteria, and bacteria inhibiting or benefiting algae,

depending on the bacteria and algae that are present in the co-culture system. In

addition, microalgae is an effective platform for promoting electron and energy flow

in a microalgal-bacteria symbiosis (Fig.1).

2. Selection of microalgae species used in wastewater treatment

Using microalgal cultivation as a tertiary wastewater treatment process started

since 1970s. The most desired characteristics of algae for the use in wastewater

treatment and biofuel production include higher growth rate, high lipid content and

productivity, higher tolerance to the possible pollutants - metal ions and toxic

compounds present in the treated wastewater, high NH4 + tolerance, high O2 generation

6
rates, high CO2 sinking capacity, and robust growth properties with improved

tolerance for varied environmental conditions. These criteria are of prime importance

and algal growth has been reported as the limiting factor in nutrient and pollutant

removal efficiency of an algae-bacteria consortium. The selection of microalgae for a

particular treatment option can be based on the knowledge about the indigenous

species in such wastewaters, making use of their characteristics for our advantage.

Several microalgal species (such as Chlorella sp. (Wang et al., 2015b), Scenedesmus

sp. or Desmodesmus sp. (Ji et al., 2014; Martinez et al., 2000), Neochloris sp. ,

Chlamydomonas sp. (Xiong et al., 2016), Nitzschia sp., and Cosmarium sp.

(Daneshvar et al., 2007)) have been applied for various types of wastewater

treatments coupled with biofuels production under sterilized or non-sterilized

conditions (Fig. 2). Among them, species of the genera Chlorella, Scenedesmus, and

some cyanobacteria are the most employed species in various wastewater treatments

due to their high growth rate, high environmental tolerance, and high lipid/starch

accumulation potential (Kim et al., 2016; Wang et al., 2015b). For instance, Chlorella

sp. is widely applied in the wastewater treatment because of its enhanced ability in

removing nitrogen, phosphorus, and chemical oxygen demand (COD), while

Scenedesmus sp. can be cultivated in high saline piggery wastewater (Kim et al., 2016)

and high COD-loading swine wastewater (Prandini et al., 2016). Zhou et al (Zhou et

al., 2011) isolated microalgal strains from various wastewater treatment sites and

found that five strains (of the genera Chlorella sp., Hindakia sp., Scenedesmus sp.,

and Auxenochlorella protothecoides) had higher biomass and lipid productivity. Of

7
those, Scenedesmus sp. had a biomass productivity of 247 mg L-1 d-1 with 30% lipids

per dry weight. In another study, when grown in municipal wastewater, a

Scenedesmus sp. strain had a biomass productivity of 132.4 mg L-1 d-1 with a lipid

content of 11.04% w/v (McGinn et al., 2012). Oil mill effluent was also used for the

culture of Scenedesmus sp. Since the effluent caused darkening of the medium,

heterotrophic growth of Scenedesmus sp. was observed with a biomass productivity of

120 mg L-1 d -1 and a lipid and carbohydrate content of 164 mg L-1 and 174 mg L-1,

respectively (Di Caprio et al., 2015).

Botyrococcus braunii, a green freshwater microalga, is known for its unique

ability to constitutively synthesize and store high quantities of a wide variety of lipids

(Tanabe et al., 2012). Chinnasamy et al used carpet industry wastewater as a nutrient

source for growing B. braunii and a biomass of 340.4 mg L-1 d-1 and a lipid

productivity of 13% DW was obtained (Chinnasamy et al., 2010). It is worth noting

that the lipid production obtained from the axenic culture of B. braunii is equivalent to

oil production associated with the consortium of five microalgal strains and the

estimated oil yield from biomass produced with untreated wastewater were 3675 L

ha-1 yr-1 and 3830 L ha-1 yr-1, respectively, for B. braunii and the consortium. Although

B. brauni can accumulate lipids higher than this quantity, the production costs offset

by using wastewater must also be considered. In a related study, untreated carpet

inudtsry wastewater was used for the cultivation of a microalgal consortium in

polybags. The resultant biomass was rich in proteins (∼53.8%) with a biomethane

generation potential of about 12,128 m3 biomethane ha-1 year-1 (Chinnasamy et al,

8
2010a). The marine microalga Nannochloropsis gaditana was grown with centrate

(i.e., concentrated municipal wastewater generated during sludge centrifuge) as a sole

nutrient source diluted in seawater (30-50%), with a biomass productivity of 0.4 g L-1

d-1 and lipid accumulation of 20-25 % DW. The biomass content was similar to that

obtained with nutrient media under laboratory conditions, indication the richness of

removable nutrients in centrate. In addition, some macroalgae, such as Sargassum

cymosum, have been used for biosorption or bio-accumulation of heavy metals (such

as Cu, Pb, and Ge) (Costa et al., 2016). Some brown-macroalgae (e.g., Ascophyllum

nodosum, Fucus spiralis, Laminaria hyperborea and Pelvetia canaliculata) can

remove transition metal ions from petrochemical wastewaters (Cechinel et al., 2016).

These observations can be used as a guideline to choose the microalgal strain for a

particular purpose, as the choice of microalgae for wastewater treatment is based

primarily on the pollutants the microalgae need to deal with. In other words, for

different wastewater sources, the mechanisms of microalgal wastewater treatment

should be different. Figure 3 illustrates the four main mechanisms (i.e., bio-adsorption,

bio-accumulation, bio-coagulation and bio-conversion) involved in microalgae-based

wastewater treatment, which will be critically discussed in the following sections.

3. Potential of algae-bacteria consortium system used for wastewater treatment

It has been observed that the co-cultivation of algae and bacteria can stimulate

algal growth; algae and bacteria are widely known to form consortia in nature. The

symbiotic association of algae and bacteria in an algae-bacteria consortium is

9
summarized in Table 1. Some researchers pointed out that even the axenic status of

laboratory microalgal cultures is questionable. Park et al (Park et al., 2008) isolated

eight bacterial strains from a laboratory stock of Chlorella ellipsoidia and found that a

single bacterium Brevundimonas sp. is capable of promoting microalgal growth. It is

believed that the associated bacteria reduce photosynthetic oxygen tension of

phototrophic microalgae by consuming O2 as their electron acceptor. It is also

postulated that the algae-bacteria symbiosis is beneficial for algae as they can supply

essential nutrients like vitamins and other compounds, as many algae are auxotrophic

for cobalamine (Croft et al., 2005). Siderophores present in some bacteria can

promote microalgal growth under iron deficient conditions (Amin et al., 2009). Also,

the extracellular sheath (composed of various sugars) present in some algae provide

sites of attachment as well as organic carbon sources for bacterial growth, and

photosynthetic oxygen as electron acceptor for aerobic respiration (Park et al., 2008).

In a similar study, it was shown that the associated bacteria plays a profound role in

improving flocculation of Chlorella vulgaris, by increasing the floc size resulting in

the sedimentation of microalgae (Lee et al., 2013). In the co-cultivation system,

microalgae-bacteria consortium is usually formed by self-coagulation between algae

and bacteria. Several advantages can be observed in an algae-bacteria consortium

system: (1) The co-cultivation of algae and bacteria not only reduces high costs, but

also decrease the spatial distance for O2 and CO2 exchange, when compared to

separate culture units. (2) Starvation, the main strategy used to trigger

carbohydrate/lipid accumulation, can be achieved by manipulating the nutrient

10
content in the medium, similar to laboratory cultivation methods. (3) Enhanced

biomass harvesting efficiency (Xu et al., 2016) because of the auto-flocculation and

co-flocculation properties of the constituent microbiota.

The formation of microalgae-bacteria consortium is dynamic and is usually

divided into four stages. Stage 1: Adsorption of microalgae onto the surface of

flocculated sludge, probably due to extracellular polymeric substance (EPS) bridging

and large superficial area of the activated sludge flocs (Salim et al., 2014). Stage 2:

Attachment of the nascent bacteria onto the surface of microalgae, mainly to the

phycosphere on the surface of algal cell and its immediate environs (Eigemann et al.,

2013). Stage 3: Uniform distribution of microalgae and bacteria in the pellet,

accompanied by growth of both of the constituent biota (Su et al., 2011). Stage 4:

Formation of mature consortium (i.e., dynamic balance between biomass attachment

and detachment), which is considered to be synergistic between algae and bacteria

(Xu et al., 2016). Algal exudates are the main carbon source for bacteria, while algae

can also benefit from bacteria as it provides CO2 and nutrients through organic matter

decomposition. Several studies have combined microalgal photosynthesis with

conventional wastewater treatment. Indirect adhesion of bacterial symbionts on the

sheath and the direct adhesion onto the algal cell surface may reduce diffusion

distance and permit rapid and efficient exchange of substrates (Park et al., 2008).

In the algae-bacteria consortium system, algae produce oxygen for nitrification

through photosynthesis, which reduces aeration demands. Removal of nitrogen can be

achieved through assimilation by biomass and nitrification/denitrification with

11
different reaction mechanisms, which is dependent on the wastewater characteristics

and reactor operating regime. Theoretically, 1 g of N input supports the production of

16.8 g of oxygen, which is sufficient to oxidize another 5.6 g of nitrogen (Karya et al.,

2013). Therefore, the ideal nitrogen flux in the co-cultivation system is that 15% of

nitrogen is taken up by algae and 85% of them are removed through the nitrification

route, demonstrating that higher percentage of nitrifying bacteria would face oxygen

deficiency because the remaining algae would not be able to produce sufficient

oxygen for full nitrification. Su et al. found that assimilation of nitrogen into biomass

can reach 61-93% of nitrogen removal in batch wastewater treatment reactors with

co-culture of algae and bacteria (Su et al., 2011). Karya et al. also found that 81-85%

of the ammonium supplement was converted to NO3- by algae-bacteria consortium

(Karya et al., 2013). Biological nitrification has been shown to have the ability of

solving certain problems, such as poor settling characteristics, during algal treatment

of high ammonium-containing wastewaters (Su et al., 2011). A two-phase photoperiod

approach (consisting of a 12 h:60 h Light-Dark cycle, followed by a 12 h:12 h

Light-Dark cycle) was capable of efficient nutrient removal and enhanced biomass

and lipid productivity by a well-balanced microbial consortia consisting of algae

(Scenedesmus) and bacteria (Flavobacteria, Sphingobacteria and Proteobacteria (Lee

et al., 2016). A similar approach of alternate light and dark photoperiod based

cultivation was used for shortcut nitrogen removal using algae-bacterium consortia,

where O2 production by photosynthetic activity stimulates

ammonium-oxidizing-bacteria (AOB) during light period, and during the dark period,

12
dissolved oxygen (DO) is quickly consumed by microbial activity and algal

respiration to promote denitritation (Wang et al., 2015a). The total energy influx into

the photobioreactors comes from both organic nutrients and light sources.

Mass balance between microalgae and bacteria could be elucidated through a

stable culture system (e.g., chemostat), while photosynthetic activity (such as the CO2

assimilation rate by algae) should be monitored in a precise and dynamic way (Lee et

al., 2016). In the microalgae-bacteria consortia, the bacterial proportion serves as the

CO2 supplier to provide inorganic carbon for algae growth, where nearly 300 bacteria

units per algae unit were proposed to ensure a stable CO2 supply. However, it would

be strongly affected by the species, bioreactors, or growth condition.

The three main algae-bacteria consortium currently in use are as follows: (1)

algae plus wastewater, (2) algae plus activated sludge, (3) co-culture of algae and

assigned bacteria (such as nitrogen-fixing bacteria Azospirillum). Major bacteria that

usually exist in initial wastewater were identified as members of the following genera

Arcobacter, Clostridium, Desulfobulbus, Flavobacterium, Acidovorax, Bacteroides,

Propionivibrio, Dechloromonas, Flexibacter, Pseudomonas, Rhodobacter,

Enhydrobacter, Sphingobacterium, Paludibacter, Smithella, Pedobacter, and

Zoogloea . The abundance of Proteobacteria, Flavobacteria and Sphingobacteria in

algal cultures has been previously reported (Sapp et al., 2007). Depending on the algal

species in the algae-bacteria consortia and the immediate environs during the various

growth stages of algae, the associated bacterial community structure can be affected in

a species-specific manner. Co-cultures of bacteria and algae have been examined,

13
including both bacteria-algae consortia in nature and artificially induced symbioses.

For natural symbiosis, commonly occurring bacterial species of Brevundimonas and

Sphingomonas has been reported (Tate et al., 2013). For artificially induced

algae-bacteria consortium, the use of nitrogen-fixing bacterium Azospirillum has been

studied (de-Bashan, 2002). De-bashan et al identified the stimulation of microalgal

growth by associated bacteria and coined the term “MGPB” – Microalgal growth

promoting bacteria. The preliminary study included co-immobilization of Chlorella sp.

and Azospirillum brazileinse for efficient removal of nutrients from wastewater

(de-Bashan et al., 2004). Currently, many studies have investigated the co-culture of

Chlorella sp. with Azospirillum sp. in alginate beads, in order to treat wastewater via

the microalgae-bacteria consortium (de-Bashan, 2002) (Table 2).

Even though there are many reports regarding treatment of various wastewaters

with microalgae-bacteria consortium, the focus was mainly on the efficiency of

nutrient removal, not biomass or lipid production. In view of this, a co-culture of

Scenedesmus sp. and indigenous municipal wastewater bacteria for the treatment of

pretreated municipal wastewater was carried out. An alternative light and dark cycle

was employed during cultivation which improved biomass productivity and lipid

content. The biomass productivity and lipid productivity were 282.6 and 71.4 mg L-1

day-1) (Yang et al., 2000). A similar co-culture of Chlorella sorokiniana and aerobic

bacteria on primarily treated potato industry wastewater with alternate light and dark

cycle yielded a biomass productivity of 26 mg L-1 day-1 with a lipid content of 30%

DW (Al-Awadhi et al., 2003). In addition, compared with photoautotrophic culture,

14
mixotrophic cultivation (e.g., algal-bacteria co-culture system) could achieve higher

growth rates because of the production of beneficial photosynthetic metabolites. For

example, under mixotrophic conditions, the carbohydrate productivity of

Scenedesmus could be increased three fold than autotrophic cultivation, due to

corresponding organic carbon addition and enhanced algal biomass productivity (Ji et

al., 2015). Moreover, extracellular organic matters (EOM) were frequently secreted by

the algae-bacteria consortium. Starvation has been confirmed to promote bacterial

aggregation and EPS stimulation (Zhou et al., 2014). EPS stimulation increases the

settle ability of biomass and reduces process costs in biomass harvesting. The

increased EOM may originate from reserve starch, which could play a crucial role in

the flocculation of biomass via its highly charged polymer structure (Mikulec et al.,

2015). However, more studies are needed to elucidate the mechanisms of EOM

secretion and the role of bacteria (promoting or limiting) in an algae-bacteria

consortium.

4. Biological treatments of different wastewaters using microalgal cultures

4.1 Pharmaceutical wastewater

Pharmaceutically active compounds (PhACs), mainly present in aquatic

environments, are emerging as a severe risk for both wildlife and humans (Kolpin et

al., 2002). Because of the flourishing pharmaceutical industry, unfortunately, many

PhACs have been systematically detected in wastewater over past decades (Ternes,

1998). Till date, the occurrence of more than 200 different PhACs in water body has

15
been reported. For instance, the antibiotic ciprofloxacin has been found in aquatic

environments at a concentration of up to 6.5 mg L-1 (Petrie et al., 2015). Harmful

concentrations of these compounds in higher trophic level organisms and long term

effects via bio-magnification in food chains are evident because of the extensive

therapeutic use of pharmaceutical compounds (PCs) (Kelly et al., 2007). Among these,

the most commonly seen drugs are non-steroidal anti-inflammatory drugs (NSAIDs)

ibuprofen and diclofenac, antibiotics (such as erythromycin, roxithromycin,

ketoconazole, quinolones, fluoroquinolones), β-blockers (propranolol),

anti-depressants and antiepileptics (carbamazepine) (Petrie et al., 2015).

Microalgae are the primary producers in aquatic food chains, and they are the

key indicators for assessing water quality and eco-toxicity of pollutants (Stevenson &

Graham, 2014). Bioremediation of contaminated waters by mixotrophic microalgae is

a solar power-driven, ecologically comprehensive and sustainable reclamation

strategy (Xiong et al., 2016). Carbamazepine (CBZ) is one of the most investigated

compounds in pharmaceutical industry based wastewater for bio-remediation by

microalgae. Chlamydomonas mexicana and Scenedesmus obliquus were evaluated for

the toxicity, cellular stress and biodegradation stability of carbamazepine (CBZ)

(Xiong et al., 2016). The growth of S. obliquus was significantly inhibited (nearly

97%) at 200 mg CBZ L-1, whereas under the same condition, only 30% inhibition was

observed for C. mexicana (Xiong et al., 2016). Moreover, C. mexicana and S.

obliquus could achieve a maximum of 35% and 28% biodegradation of CBZ,

respectively (Xiong et al., 2016). It was also found that a mixture of pharmaceutical

16
compounds (including CBZ) could strongly decrease the activity of ATP synthase in

Pseudokirchneriella subcapitata (Xiong et al., 2016), suggesting that PhACs can

interfere with energy transduction in the mitochondria and chloroplast of algae

(Vannini et al., 2011). Recently, Matamoros et al. (Matamoros et al., 2016) reported

that the microalgal consortium, containing Chlorella sp. and Scenedesmus sp. can

successfully remove 20% of CBZ from urban and synthetic wastewater, as shown in

Table 3.

The removal of pharmaceutical compounds from wastewater by algae-bacteria

consortium could be by means of bioaccumulation or biodegradation. However, the

elucidation of the actual mechanism and the role of bacteria in such processes needs

to be further investigated. A study regarding CBZ removal by microalgae showed that

some key enzymes (e.g., SOD, and CAT) of phototrophic microorganisms would

protect against the reactive oxygen species (ROS) toxicity via regulation of

anti-oxidative defense mechanisms (Xiong et al., 2016), which are commonly

identified as the biomarkers (Zhang et al., 2012) (Fig. 3). The superoxide dismutase

(SOD) activity in C. mexicana was increased by 1.3-fold at low CBZ concentrations

(50 mg L-1) whereas the SOD activity at higher concentrations (200 mg L-1) was

significantly decreased (Xiong et al., 2016). In another related study, Zhang et al.

(Zhang et al., 2012) found that low CBZ concentration (0.5 to 10 mg L-1) could

induce superoxide dismutase (SOD) and catalase (CAT) activities in microalgal cells.

Freitas et al. (Freitas et al., 2015) also demonstrated that the SOD activity in

Scrobicularia plana and Diopatra neapolitana would increase at the extremely low

17
CBZ concentration (<6 µg L-1), whereas the SOD activity of both species decreased

with an increase in CBZ concentrations. This is because the stress induced by CBZ

would significantly enhance the generation of ROS in algal cells, resulting in

membrane lipid peroxidation and functional damage (Xiong et al., 2016). In addition,

the ROS formation caused by accumulated CBZ could damage the cell structure and

related physiological and biochemical processes (Dordio et al., 2011; Ke et al., 2010).

It has also been reported that the microalgal biodegradation prevented the formation

of toxic intermediates (such as carcinogenic compounds), which can be considered a

significant achievement for the microalgal biodegradation (Xiong et al., 2016).

Considering the fact that thousands of PhACs and tons of PhACs containing

wastewater are generated every year, it is of urgent demand to find effective means to

deal with this type of wastewater. A promising area for future research is to identify

the role of algae-bacteria consortium in degradation of PhACs, thereby providing a

feasible way for efficient treatment of such wastewaters.

4.2 Dye-containing wastewater

Textile industries commonly release a vast amount of wastewater, which contains

various fabric dyes as the main constituent. There are more than 100,000

commercially available dyes, and more than 7 × 105 tons of dyestuffs are produced

worldwide annually (Robinson et al., 2001). Textile industrial wastewater is

characterized by strong color, high salinity, high temperature, variable pH and high

chemical oxygen demand (COD). In general, most textile industries handle a huge

18
quantity of synthetic dyes, sodium sulphide, Glauber salt (in dye bath solution), and

hydrogen peroxide (as oxidizing agent) (Vijayaraghavan & Shanthakumar, 2015).

Algae have been used to remove these colorful dyes through biosorption or reductive

mechanisms (Fig. 3). It was reported that more than 30 azo dyes could be decomposed

by Chlorella sp. and Oscillatoria sp. to aromatic amines that were further completely

metabolized by bacterial cultures (Jinqi & Houtian, 1992). Marungrueng et al.

(Marungrueng & Pavasant, 2007) also mentioned that the Caulerpa lentillifera can

effectively remove three basic dyes (CI Basic Blue, CI Basic Red, CI Basic Blue) by

biosorption.

To use microalgal cells as the bio-sorbent or bio-coagulator, their growth

tolerance to various dyes should be considered. For instance, studies have shown that

C. vulgaris has high tolerance to azo dyes, and it can even grow in the presence of 400

mg L-1 of tectilon yellow 2G (a mono-azo dye) (Acuner & Dilek, 2004) (Table 4). In

comparison, some algae species (e.g., Pseudokirchneriella subcapitata, Selenastrum

caprincornutum) are very sensitive to azo dyes where their IC50 can be as low as 0.5

mg L-1 (Novotny et al., 2006). C. vulgaris can remove 63-69% of color from tectilon

yellow 2G by converting it to aniline (Acuner & Dilek, 2004). The ability of C.

vulgaris to break down the azo bond has also been observed by Lim et al (Lim et al.,

2010) while treating textile wastewater with the microalgal strain. Macroalgae, such

as Caulerpa lentillifera (a green seaweed), can also remove basic dyes via biosorption

and the dye adsorption efficiency of this green seaweed was highly correlated with the

dye concentration in wastewater (Marungrueng & Pavasant, 2006). On the other hand,

19
cyanobacteria, such as Synechocystis sp. and Phormidium sp., was also found to

effectively remove reactive dye metabolically and the removal efficiency was

enhanced with the addition of specific plant growth regulator triacontanol hormone

(Karacakaya et al., 2009).

Previous studies showed that the mechanism of color removal by Chlorella sp. is

mainly due to biosorption (Chu et al., 2009) as shown in Fig. 3. To further understand

the mechanism of microalgal biosorption, the related adsorption equilibriums were

applied for evaluating the adsorption type. Freundlich, Langmuir, Redlich-Peterson

and Koble-Corrigan adsorption models were employed for the mathematical

description of biosorption equilibrium, together with evaluating the isotherm

constants at different temperatures (Aksu & Tezer, 2005). Algae have been found to

be potential bio-sorbents due to their rapid growth, relatively high surface area, and

high binding affinity. Algal cell walls offer a host of functional groups including

amino, carboxyl, sulphates, phosphates, and imidazoles associated with

polysaccharides, alginic acid and proteins for binding various pollutants. The

adsorption of reactive dyes (e.g., Remazol Red and Remazol Golden Yellow) by dried

biomass of C. vulgaris followed the Langmuir model (Aksu & Tezer, 2005),

suggesting that the dyes were adsorbed on monolayer coverage on homogenous sites

of algal cell wall. On the other hand, conformation to Freundlich model suggests the

existence of a heterogeneous surface with sorption sites of different affinities. The

possibility of the dyes being absorbed into the cell and then transformed through

bioconversion is worthwhile for further investigation.

20
Several studies have provided some exciting ideas for algae-based dye

wastewater treatment. First, extracellular polymeric substances (known as EPS) and

fluidity of the membrane could contribute to the adsorption process. For instance,

Enteromorpha polysaccharide is one of the common fouling generated by green algae,

which has caused serious environmental problems in past years. However, these

Enteromorpha polysaccharides now are serving as a new-type coagulant for color

removal because of its possible role as a bridging agent in floc formation (Zhao et al.,

2014). Second, the biofilm produced by algae-bacteria consortium system has high

potential for dye wastewater treatment. A novel photobioelectrochemical system

(PBES) was developed by acclimating algae-bacteria biofilm in anode and cathode,

using C. vulgaris and indigenous wastewater bacteria, respectively, as the inoculum

(Sun et al., 2015). The synergy between C. vulgaris and mixed bacteria was

responsible for the successful operation of PBES, which can be potentially applied to

treat wastewater containing azo dye, along with additional benefits of enhanced azo

dye degradation, high net power output, and buffer minimization (Sun et al., 2015).

Third, cyanobacteria have also been applied for dye pollutants removal because of its

coagulation behavior and floc characteristics (i.e. hydroxide flocs) (Karacakaya et al.,

2009). Finally, different algal species were compared for various dye removal

efficiencies to provide the detailed information on algae-based dye wastewater

treatments. In particular, the ability of C. vulgaris, Lyngbya lagerlerimi, Nostoc lincki,

Oscillatoria rubescens, Elkatothrix viridis and Volvox aureus to decolorize and

remove methyl red, orange II, G-Red (FN-3G), basic cationic, and basic fuchsin was

21
investigated (El-Sheekh et al., 2009). The color removal efficiency of these algae is

dependent on different species, growth stage, cultivation conditions, as well as the

molecular structure of the dye.

4.3 Metal-containing wastewater

Many studies have clearly demonstrated the potential of metal removal from

wastewater by algal biomass. Cell walls of algae and cyanobacteria are composed of

carbohydrates and polysaccharides with negatively-charged groups (e.g. hydroxyl,

amino, carboxyl or sulfhydryl group). Most positively-charged metals can tightly bind

to the negatively-charged ligand groups, which is the basis of metal removal from

metal containing wastewater (Fig. 3). In addition to adsorption onto cell surfaces,

other mechanisms for efficient metal removal include bioaccumulation (uptake by

cells or incorporation into vacuoles), formation of chelates (such as aragonite

structures) and internal or surface precipitation. Thus, algae growing in wastewaters

may provide a simple, long-term strategy for removal of metal pollutants. Meanwhile,

dried algal, cyanobacterial and bacterial biomass can also be applied for removal of

heavy metals from wastewaters (such as petrochemical wastewater) as cation

exchanger, and metals can be recovered subsequently by desorption with acids or

other desorbing agents (Vijayaraghavan & Yun, 2008). Efficient removal of copper

(80%) and cadmium (100%) from metal-containing wastewater via a mixture of dried

algal biomass was observed obtaining a maximum removal rate within 5 minutes

contact time (Loutseti et al., 2009)). However, it was also reported that heavy metals

22
are potent inhibitors of photosynthesis because of their ability to replace or block the

prosthetic metal atoms in the active sites of certain photosynthesis related enzymes.

On the contrary, the acidic functional groups on bacterial cell walls or extracellular

polysaccharides produced by bacteria can bind significant concentrations of aqueous

cations (e.g., Cd), which in turn influences the speciation, distribution and mobility of

such cations in aqueous systems (Ueshima et al., 2008). An algae-bacteria consortium

comprising C. sorokiniana and R. basilensis was found capable of metabolizing

salicylate with subsequent removal of heavy metals from solution (Munoz et al.,

2006). It was also reported that copper cations could be removed more efficiently by

an algae-bacteria consortium compared with individual constituent organism (Boivin

et al., 2007), suggesting that metal removal by the consortium is a synergistic effect.

Future research direction for algae-bacteria consortium based heavy metal

removal can involve the following points to be addressed: (1) The mechanism of

algae-bacteria consortium for heavy metal removal, such as biosorption,

bio-convention or bio-accumulation, should be systematically investigated, (2) The

synergistic effects of algae and bacteria need more elucidation. (3) Ecological

cation-exchanger can be developed, since dried biomass of algae and bacteria possess

the ability of ion exchange due to charged cell walls.

4.4 Agro-industrial wastewater

Agro-industrial wastewaters include potato processing wastewater, swine

wastewater (Wang et al., 2015b), livestock wastewater, dairy wastewater(Gentili,

23
2014), slaughterhouse wastewater , fish farm wastewater and some digestion effluent.

One common characteristic of agro-industrial wastewater is high ammonium

concentration, which is highly correlated to eutrophication (Wang et al., 2015b).

Many microalgae can grow well under “nutrient-rich” environment and rapidly

convert the nutrients (e.g., ammonium used as a nitrogen source) contained in

agro-industrial wastewaters to biomass (Wang et al., 2015b). The microalgae

Neochloris oleoabundans, C. vulgaris and S. obliquus were cultivated using

agro-zootechnical digestate and compared for their ability in nitrogen removal and

lipid production (Franchino et al., 2013). However, some agro-industrial wastewaters

are characteristic of high COD levels, (i.e., 20,180 mg L-1) (Wang et al., 2015b), and

thus it may not be easy to directly treat the original wastewaters by microalgae. Thus,

most studies on microalgae-based treatment of such wastewaters are conducted with

appropriate dilution of original wastewater or employing anaerobic digestion as the

pretreatment. Under optimal conditions, the ammonium removal efficiency by

microalgae can often reach higher than 90% (Wang et al., 2012; Wang et al., 2015b).

Cyanobacteria are also very efficient in nitrogen removal. Some cyanobacteria, such

as Oscilatoria sp., Anabaena sp. and Spirulina sp., are capable of utilizing elemental

nitrogen as their sole nitrogen source, by the reduction of N2 to ammonium (Markou

& Georgakakis, 2011). In general, the order in which cyanobacteria prefer to utilize

nitrogen is ammonium>NO3->N2. The uptake of nitrate is mostly light dependent, and

since the reduction of nitrite consumes energy, cyanobacteria are inclined to assimilate

the reduced nitrogen (e.g., ammonium). Thus, co-culture of microalgae and

24
cyanobacteria for ammonium-rich agro-industrial wastewater would be a potential

solution. Moreover, it has been reported that nitrate consumption in wastewater will

be reduced due to the high turbidity, which will increase treatment costs. In addition,

high concentrations of ammonium will inhibit uptake of nitrate, because ammonium

represses the synthesis of nitrate reductase, while high nitrate concentration inhibits

ammonia uptake. High temperatures favor the formation of free ammonia. Although

free ammonia is generally toxic to photosynthetic organisms, the toxicity appears to

be reduced in alkalophilic species such as S. platensis (Belkin & Boussiba, 1991).

5. Photobioreactor design for microalgae-based wastewater treatment

Currently, most of the microalgal bioreactors can be categorized into open and

closed systems with suspended or immobilized cell cultures. Examples of

photobioreactors used for wastewater treatment with microalgae-bacteria consortium

are summarized in Table 5. The open pond culture system is generally regarded as the

most economical process among algae cultivation systems because of low operating

costs and simple operation. However, open pound systems depend a lot on the natural

environment of pond site, and contamination is liable to happen. Types of open pond

system that are currently in use include slope system, raceway ponds and circular

ponds and these have been developed for decades. Hernández et al. used a high rate

algal pond (HRAP) for treatment of slaughterhouse wastewater and showed >70%

removal of ammonium under all experimental conditions by an algae-bacteria

consortium. In addition, oxidation of both organic matter and ammonium was

25
achieved by the C. sorokiniana-mixed bacterial culture from the activated sludge

process in the tubular biofilm photobioreactor (Gonzalez et al., 2008).

Using closed photobioreactors for microalgal cultivation usually require higher

capital and operating costs but they are in general more efficient in cell growth due to

better control of the culture conditions and lower contamination risks. Biomass

production and nutrient removal efficiencies can be improved by efficient culture

techniques. In recent years, there are increasing interests in utilizing immobilized

cultures (in particular, biofilm reactors) for microalgae cultivation to obtain better

performance during microalgae-based wastewater treatment in contrast to

conventional suspended cultures. Biofilm based culturing is the common immobilized

culture system, which is promising for increasing the algal culture density with less

water source and land space (Robinson et al., 2001). It has obvious advantages for

biomass and lipid/carbohydrate production, nutrient removal and lower energy cost

compared with suspended culture photobioreactors (e.g., tubular photobioreactor,

plate photobioreactor). Christenson and Sims (Christenson & Sims, 2012) compared

the treatment of secondary effluent of municipal wastewater by rotating algae biofilm

reactor (RABR), with the highest biomass productivity of 31 g m-2 d-1, which is much

higher than suspended system (biomass productivity of 7.4 g m-2 d -1). The EOM

secreted by microalgae consists of polysaccharides, proteins, nucleic acids, and

phospholipids, which contribute to biofilm formation as well (Wingender et al., 2012).

Typical microalgal biofilm reactors employed in wastewater treatment are

limited to horizontal biofilm reactor and rotating algal biofilm reactor (RABR). The

26
horizontal biofilm reactor can receive light energy effectively, however, a large

surface area is needed in order to assure biomass productivity and nutrient removal. In

most studies, the horizontal biofilm photobioreactors attained higher nutrient removal

but lower algal biomass production. Improvement of photobioreactor design can

effectively promote nutrient removal. Wu et al. (Wu et al., 2011) proposed a hybrid

bioreactor which simultaneously supports heterotrophic and autotrophic algal-bacteria

consortium for high-loading nutrients removal, with 81% total phosphorus and 86%

ammonium removal efficiency. In addition, common wastewater treatment reactor

(e.g., sequencing batch reactor, PBR) can be used for algae cultivation as well (Van

Den Hende et al., 2014). The effective alternate light-dark cultivation of

microalgae-bacteria consortium in a photo-sequencing batch reactor (PSBR) can

accomplish efficient N and P removal without aeration (Wang et al., 2015a). Some

biofilm-based wastewater treatment tank can also be used for microalgae wastewater

treatment, such as algae membrane bioreactor (A-MBR). Xu et al used an A-MBR to

cultivate Chlorella vulgaris in continuous mode, obtaining the highest total N (TN)

removal efficiency of 73.4 ± 6.3% and total P (TP) removal efficiency of 91.3 ± 3.8%

at a solid retention time (SRT) of 10 d and a HRT of 24 h (Xu et al., 2015). It was

found that a shorter SRT seemed to favor biomass production, while a longer SRT led

to poorer cell growth. According to Xu et al.'s study, the highest algal productivity

(131.7 g m-3/d or 22.4 g m-2/d) was observed at the a short SRT of 5 d (Xu et al.,

2015). In addition, the operation of a microalgal biofilm photobioreactor would be

helpful for increasing nutrient removal efficiency. Fernadez et al developed a

27
hierarchical control strategy for microalgal production in a tubular photobioreactor,

which efficiently reduced the algal cultivation cost to around 100€/kg and also

lowered the CO2 loss to 167 g/d (Fernández et al., 2016).

In some microalgae-bacteria consortium biofilm reactors, mechanical aeration

can be replaced by photosynthetic aeration and EPS can serve as flocs (Van Den

Hende et al., 2014). The stability of the flocculated microalgal-bacterial biomass was

significantly affected by extracelluler polymer substances (EPS). Moreover, the

bacterial exopolymer can improve aggregation possibilities of algae-bacteria

consortium for increasing sedimentation efficiency of the biomass. It can also be used

as stabilizers for floc formation.

In previous Using carrier materials to immobilize the microalgal-bacterial culture

is another method for improving nutrient removal. However, the price of such carrier

materials must be very low to minimize the cost associated with wastewater treatment.

The carrier materials should also be non-toxic (environmentally friendly), structurally

stable, promote enhanced mass transfer efficiency and economic. Muñoz et al.

(Munoz et al., 2009) used a biofilm photobioreactor for C. sorokiniana-R. basilensis

consortium immobilized onto foamed-glass beads carrier and onto reactor wall for

treating salicylate contaminant in wastewater. In addition, algae immobilized with

active sludge for simultaneous COD and nutrients removal has attracted more

attention recently. In addition to better waste reduction performance with the

immobilized-cell system, it may also be easier and less expensive to collector harvest

the microalgal cells from biofilms for further uses, making it an extra benefit by using

28
the immobilized-cell culture. Considering the lack of information regarding

immobilized algae-bacteria systems, more carriers/supports should be examined for

the immobilization of algae-bacteria consortium in future studies.

6. Conclusions

Microalgae and microalgae-bacteria consortium can be successfully applied for

the treatment of nutrients-rich wastewater, such as agro-industrial wastewater,

municipal wastewater, pharmaceutical wastewater and textile dye wastewaters by

biosorption or bio-conversion pathways. Several advantages are clearly observed in an

algae-bacteria consortium system. The resulting microalgal biomass obtained from

wastewater treatment can be used for biofuels production or other applications to gain

additional benefits, making it a feasible and reliable process for dual purposes of

waste reduction and biofuels generation. Nevertheless, some mechanisms involved in

pollutant removal and bio-conversion in microalgae-based wastewater treatment are

still not clearly understood and require further investigations.

Acknowledgments

This work was supported by Open Project of State Key Laboratory of Urban

Water Resource and Environment, Harbin Institute of Technology (HCK201607). The

authors also gratefully acknowledge the support received from Taiwan’s Ministry of

Science and Technology (MOST) under grant numbers 105-3113-E-006-003,

104-2221-E-006 -227-MY3, and 103-2221-E-006 -190 -MY3.

29
References

1. Abed, R.M.M., Koster, J. 2005. The direct role of aerobic heterotrophic bacteria associated

with cyanobacteria in the degradation of oil compounds. International Biodeterioration &

Biodegradation, 55(1), 29-37.

2. Acuner, E., Dilek, F.B. 2004. Treatment of tectilon yellow 2G by Chlorella vulgaris. Process

Biochemistry, 39(5), 623-631.

3. Aksu, Z., Tezer, S. 2005. Biosorption of reactive dyes on the green alga Chlorella vulgaris.

Process Biochemistry, 40(3-4), 1347-1361.

4. Al-Awadhi, H., Al-Hasan, R.H., Sorkhoh, N.A., Salamah, S., Radwan, S.S. 2003.

Establishing oil-degrading biofilms on gravel particles and glass plates. International

Biodeterioration & Biodegradation, 51(3), 181-185.

5. Amin, S.A., Green, D.H., Hart, M.C., Kupper, F.C., Sunda, W.G., Carrano, C.J. 2009.

Photolysis of iron-siderophore chelates promotes bacterial-algal mutualism. Proc Natl Acad

Sci U S A, 106(40), 17071-17076.

6. Belkin, S., Boussiba, S. 1991. Resistance of Spirulina-Platensis to Ammonia at High Ph

Values. Plant and Cell Physiology, 32(7), 953-958.

7. Boelee, N.C., Temmink, H., Janssen, M., Buisman, C.J., Wijffels, R.H. 2011. Nitrogen and

phosphorus removal from municipal wastewater effluent using microalgal biofilms. Water

Res, 45(18), 5925-5933.

8. Boivin, M.E., Greve, G.D., Garcia-Meza, J.V., Massieux, B., Sprenger, W., Kraak, M.H.,

Breure, A.M., Rutgers, M., Admiraal, W. 2007. Algal-bacterial interactions in metal

contaminated floodplain sediments. Environ Pollut, 145(3), 884-894.

9. Cechinel, M.A.P., Mayer, D.A., Pozdniakova, T.A., Mazur, L.P., Boaventura, R.A.R., de

Souza, A.A.U., de Souza, S.M.A.G.U., Vilar, V.J.P. 2016. Removal of metal ions from a

petrochemical wastewater using brown macro-algae as natural cation-exchangers. Chemical

Engineering Journal, 286, 1-15.

10. Cheriaa, J., Bettaieb, F., Denden, I., Bakhrouf, A. 2009. Characterization of new algae

isolated from textile wastewater plant. Journal of Food Agriculture & Environment, 7(3-4),

700-704.

30
11. 11. Chinnasamy, S., Bhatnagar, A., Hunt, R. W., Das, K.C. 2010. Microalgae cultivation in a

wastewater dominated by carpet mill effluents for biofuel applications. Bioresour Technol,

101, 3097–3105

12. Chinnasamy, S., Bhatnagar, A., Claxton, R., Das, K.C. 2010a. Biomass and bioenergy

production potential of microalgae consortium in open and closed bioreactors using

untreated carpet industry effluent as growth medium. Bioresour Technol, 101(17),

6751-6760.

13. Christenson, L.B., Sims, R.C. 2012. Rotating algal biofilm reactor and spool harvester for

wastewater treatment with biofuels by-products. Biotechnol Bioeng, 109(7), 1674-1684.

14. Chu, W.L., See, Y.C., Phang, S.M. 2009. Use of immobilised Chlorella vulgaris for the

removal of colour from textile dyes. Journal of Applied Phycology, 21(6), 641-648.

15. Costa, G.B., de Felix, M.R., Simioni, C., Ramlov, F., Oliveira, E.R., Pereira, D.T., Maraschin,

M., Chow, F., Horta, P.A., Lalau, C.M., da Costa, C.H., Matias, W.G., Bouzon, Z.L., Schmidt,

E.C. 2016. Effects of copper and lead exposure on the ecophysiology of the brown seaweed

Sargassum cymosum. Protoplasma, 253(1), 111-125.

16. Croft, M.T., Lawrence, A.D., Raux-Deery, E., Warren, M.J., Smith, A.G. 2005. Algae acquire

vitamin B12 through a symbiotic relationship with bacteria. Nature, 438(7064), 90-93.

17. Daneshvar, N., Ayazloo, M., Khataee, A.R., Pourhassan, M. 2007. Biological decolorization

of dye solution containing Malachite Green by microalgae Cosmarium sp. Bioresour Technol,

98(6), 1176-1182.

18. de-Bashan, L. 2002. Removal of ammonium and phosphorus ions from synthetic wastewater

by the microalgae Chlorella vulgaris coimmobilized in alginate beads with the microalgae

growth-promoting bacterium Azospirillum brasilense. Water Research, 36(12), 2941-2948.

19. de-Bashan, L.E., Hernandez, J.P., Morey, T., Bashan, Y. 2004. Microalgae growth-promoting

bacteria as "helpers" for microalgae: a novel approach for removing ammonium and

phosphorus from municipal wastewater. Water Res, 38(2), 466-474.

20. de-Bashan, L.E., Trejo, A., Huss, V.A., Hernandez, J.P., Bashan, Y. 2008. Chlorella

sorokiniana UTEX 2805, a heat and intense, sunlight-tolerant microalga with potential for

removing ammonium from wastewater. Bioresour Technol, 99(11), 4980-4989.

21. Di Caprio, F., Altimari, P., Pagnanelli, F. 2015. Integrated biomass production and

31
biodegradation of olive mill wastewater by cultivation of Scenedesmus sp. Algal Research, 9,

306-311.

22. Dordio, A.V., Belo, M., Teixeira, D.M., Carvalho, A.J.P., Dias, C.M.B., Pico, Y., Pinto, A.P.

2011. Evaluation of carbamazepine uptake and metabolization by Typha spp., a plant with

potential use in phytotreatment. Bioresource Technol, 102(17), 7827-7834.

23. Eigemann, F., Hilt, S., Salka, I., Grossart, H.P. 2013. Bacterial community composition

associated with freshwater algae: species specificity vs. dependency on environmental

conditions and source community. FEMS Microbiol Ecol, 83(3), 650-663.

24. El-Sheekh, M.M., Gharieb, M.M., Abou-El-Souod, G.W. 2009. Biodegradation of dyes by

some green algae and cyanobacteria. International Biodeterioration & Biodegradation, 63(6),

699-704.

25. Ertugrul, S., Bakir, M., Donmez, G. 2008. Treatment of dye-rich wastewater by an

immobilized thermophilic cyanobacterial strain: Phormidium sp. Ecological Engineering,

32(3), 244-248.

26. Fernández, I., Berenguel, M., Guzmán, J.L., Acién, F.G., de Andrade, G.A., Pagano, D.J.

2016. Hierarchical control for microalgae biomass production in photobiorreactors. Control

Engineering Practice, 54, 246-255.

27. Franchino, M., Comino, E., Bona, F., Riggio, V.A. 2013. Growth of three microalgae strains

and nutrient removal from an agro-zootechnical digestate. Chemosphere, 92(6), 738-744.

28. Freitas, R., Almeida, A., Calisto, V., Velez, C., Moreira, A., Schneider, R.J., Esteves, V.I.,

Wrona, F.J., Soares, A.M.V.M., Figueira, E. 2015. How life history influences the responses

of the clam Scrobicularia plana to the combined impacts of carbamazepine and pH decrease.

Environmental Pollution, 202, 205-214.

29. Gentili, F.G. 2014. Microalgal biomass and lipid production in mixed municipal, dairy, pulp

and paper wastewater together with added flue gases. Bioresour Technol, 169, 27-32.

30. Gonzalez, C., Marciniak, J., Villaverde, S., Leon, C., Garcia, P.A., Munoz, R. 2008. Efficient

nutrient removal from swine manure in a tubular biofilm photo-bioreactor using

algae-bacteria consortia. Water Science and Technology, 58(1), 95-102.

31. Guieysse, B., Borde, X., Munoz, R., Hatti-Kaul, R., Nugier-Chauvin, C., Patin, H.,

Mattiasson, B. 2002. Influence of the initial composition of algal-bacterial microcosms on

32
the degradation of salicylate in a fed-batch culture. Biotechnology Letters, 24(7), 531-538.

32. He, P.J., Mao, B., Shen, C.M., Shao, L.M., Lee, D.J., Chang, J.S. 2013. Cultivation of

Chlorella vulgaris on wastewater containing high levels of ammonia for biodiesel production.

Bioresour Technol, 129, 177-181.

33. Hernandez, D., Riano, B., Coca, M., Garcia-Gonzalez, M.C. 2013. Treatment of

agro-industrial wastewater using microalgae-bacteria consortium combined with anaerobic

digestion of the produced biomass. Bioresource Technol, 135, 598-603.

34. Hernandez, J.P., de-Bashan, L.E., Bashan, Y. 2006. Starvation enhances phosphorus removal

from wastewater by the microalga Chlorella spp. co-immobilized with Azospirillum

brasilense. Enzyme and Microbial Technology, 38(1-2), 190-198.

35. Ji, F., Liu, Y., Hao, R., Li, G., Zhou, Y., Dong, R. 2014. Biomass production and nutrients

removal by a new microalgae strain Desmodesmus sp. in anaerobic digestion wastewater.

Bioresour Technol, 161, 200-207.

36. Ji, M.K., Yun, H.S., Park, Y.T., Kabra, A.N., Oh, I.H., Choi, J. 2015. Mixotrophic cultivation

of a microalga Scenedesmus obliquus in municipal wastewater supplemented with food

wastewater and flue gas CO2 for biomass production. J Environ Manage, 159, 115-120.

37. Jinqi, L., Houtian, L. 1992. Degradation of azo dyes by algae. Environ Pollut, 75(3),

273-278.

38. Karacakaya, P., Kilic, N.K., Duygu, E., Donmez, G. 2009. Stimulation of reactive dye

removal by cyanobacteria in media containing triacontanol hormone. J Hazard Mater,

172(2-3), 1635-1639.

39. Karya, N.G., van der Steen, N.P., Lens, P.N. 2013. Photo-oxygenation to support nitrification

in an algal-bacterial consortium treating artificial wastewater. Bioresour Technol, 134,

244-250.

40. Ke, L., Luo, L.J., Wang, P., Luan, T.G., Tam, N.F.Y. 2010. Effects of metals on biosorption

and biodegradation of mixed polycyclic aromatic hydrocarbons by a freshwater green alga

Selenastrum capricornutum. Bioresource Technol, 101(18), 6950-6961.

41. Kelly, B.C., Ikonomou, M.G., Blair, J.D., Morin, A.E., Gobas, F.A. 2007. Food web-specific

biomagnification of persistent organic pollutants. Science, 317(5835), 236-239.

42. Khalaf, M.A. 2008. Biosorption of reactive dye from textile wastewater by non-viable

33
biomass of Aspergillus niger and Spirogyra sp. Bioresour Technol, 99(14), 6631-6634.

43. Kim, H.C., Choi, W.J., Chae, A.N., Park, J., Kim, H.J., Song, K.G. 2016. Evaluating

integrated strategies for robust treatment of high saline piggery wastewater. Water Res, 89,

222-231.

44. Kolpin, D.W., Furlong, E.T., Meyer, M.T., Thurman, E.M., Zaugg, S.D., Barber, L.B.,

Buxton, H.T. 2002. Pharmaceuticals, hormones, and other organic wastewater contaminants

in US streams, 1999-2000: A national reconnaissance. Environmental science & technology,

36(6), 1202-1211.

45. Krustok, I., Odlare, M., Shabiimam, M.A., Truu, J., Truu, M., Ligi, T., Nehrenheim, E. 2015.

Characterization of algal and microbial community growth in a wastewater treating batch

photo-bioreactor inoculated with lake water. Algal Research-Biomass Biofuels and

Bioproducts, 11, 421-427.

46. Lee, C.S., Oh, H.S., Oh, H.M., Kim, H.S., Ahn, C.Y. 2016. Two-phase photoperiodic

cultivation of algal-bacterial consortia for high biomass production and efficient nutrient

removal from municipal wastewater. Bioresour Technol, 200, 867-875.

47. Lee, J., Cho, D.H., Ramanan, R., Kim, B.H., Oh, H.M., Kim, H.S. 2013.

Microalgae-associated bacteria play a key role in the flocculation of Chlorella vulgaris.

Bioresour Technol, 131, 195-201.

48. Lim, S.L., Chu, W.L., Phang, S.M. 2010. Use of Chlorella vulgaris for bioremediation of

textile wastewater. Bioresource Technol, 101(19), 7314-7322.

49. Loutseti, S., Danielidis, D.B., Economou-Amilli, A., Katsaros, C., Santas, R., Santas, P. 2009.

The application of a micro-algal/bacterial biofilter for the detoxification of copper and

cadmium metal wastes. Bioresour Technol, 100(7), 2099-2105.

50. Mackulak, T., Mosny, M., Grabic, R., Golovko, O., Koba, O., Birosova, L. 2015. Fenton-like

reaction: a possible way to efficiently remove illicit drugs and pharmaceuticals from

wastewater. Environ Toxicol Pharmacol, 39(2), 483-488.

51. Markou, G., Georgakakis, D. 2011. Cultivation of filamentous cyanobacteria (blue-green

algae) in agro-industrial wastes and wastewaters: A review. Applied Energy, 88(10),

3389-3401.

52. Martinez, M.E., Sanchez, S., Jimenez, J.M., El Yousfi, F., Munoz, L. 2000. Nitrogen and

34
phosphorus removal from urban wastewater by the microalga Scenedesmus obliquus.

Bioresource Technol, 73(3), 263-272.

53. Marungrueng, K., Pavasant, P. 2007. High performance biosorbent (Caulerpa lentillifera) for

basic dye removal. Bioresour Technol, 98(8), 1567-1572.

54. Marungrueng, K., Pavasant, P. 2006. Removal of basic dye (Astrazon Blue FGRL) using

macroalga Caulerpa lentillifera. J Environ Manage, 78(3), 268-274.

55. Matamoros, V., Uggetti, E., Garcia, J., Bayona, J.M. 2016. Assessment of the mechanisms

involved in the removal of emerging contaminants by microalgae from wastewater: a

laboratory scale study. J Hazard Mater, 301, 197-205.

56. McGinn, P.J., Dickinson, K.E., Park, K.C., Whitney, C.G., MacQuarrie, S.P., Black, F.J.,

Frigon, J.-C., Guiot, S.R., O'Leary, S.J.B. 2012. Assessment of the bioenergy and

bioremediation potentials of the microalga Scenedesmus sp. AMDD cultivated in municipal

wastewater effluent in batch and continuous mode. Algal Research, 1(2), 155-165.

57. Mikulec, J., Polakovicova, G., Cvengros, J. 2015. Flocculation Using Polyacrylamide

Polymers for Fresh Microalgae. Chemical Engineering & Technology, 38(4), 595-601.

58. Munoz, R., Alvarez, M.T., Munoz, A., Terrazas, E., Guieysse, B., Mattiasson, B. 2006.

Sequential removal of heavy metals ions and organic pollutants using an algal-bacterial

consortium. Chemosphere, 63(6), 903-911.

59. Munoz, R., Kollner, C., Guieysse, B. 2009. Biofilm photobioreactors for the treatment of

industrial wastewaters. J Hazard Mater, 161(1), 29-34.

60. Novotny, C., Dias, N., Kapanen, A., Malachova, K., Vandrovcova, M., Itavaara, M., Lima, N.

2006. Comparative use of bacterial, algal and protozoan tests to study toxicity of azo- and

anthraquinone dyes. Chemosphere, 63(9), 1436-1442.

61. Park, Y., Je, K.W., Lee, K., Jung, S.E., Choi, T.J. 2008. Growth promotion of Chlorella

ellipsoidea by co-inoculation with Brevundimonas sp isolated from the microalga.

Hydrobiologia, 598, 219-228.

62. Petrie, B., Barden, R., Kasprzyk-Hordern, B. 2015. A review on emerging contaminants in

wastewaters and the environment: current knowledge, understudied areas and

recommendations for future monitoring. Water Res, 72, 3-27.

63. Polishchuk, A., Valev, D., Tarvainen, M., Mishra, S., Kinnunen, V., Antal, T., Yang, B.,

35
Rintala, J., Tyystjarvi, E. 2015. Cultivation of Nannochloropsis for eicosapentaenoic acid

production in wastewaters of pulp and paper industry. Bioresour Technol, 193, 469-476.

64. Prandini, J.M., da Silva, M.L., Mezzari, M.P., Pirolli, M., Michelon, W., Soares, H.M. 2016.

Enhancement of nutrient removal from swine wastewater digestate coupled to biogas

purification by microalgae Scenedesmus spp. Bioresour Technol, 202, 67-75.

65. Robinson, T., McMullan, G., Marchant, R., Nigam, P. 2001. Remediation of dyes in textile

effluent: a critical review on current treatment technologies with a proposed alternative.

Bioresource Technol, 77(3), 247-255.

66. Ruiz-Martinez, A., Serralta, J., Seco, A., Ferrer, J. 2015. Effect of temperature on ammonium

removal in Scenedesmus sp. Bioresour Technol, 191, 346-349.

67. Salim, S., Kosterink, N.R., Tchetkoua Wacka, N.D., Vermue, M.H., Wijffels, R.H. 2014.

Mechanism behind autoflocculation of unicellular green microalgae Ettlia texensis. J

Biotechnol, 174, 34-38.

68. Sapp, M., Schwaderer, A.S., Wiltshire, K.H., Hoppe, H.G., Gerdts, G., Wichels, A. 2007.

Species-specific bacterial communities in the phycosphere of microalgae? Microb Ecol,

53(4), 683-699.

69. Stevenson, J., Graham, L. 2014. Ecological assessments with algae: a review and synthesis.

Journal of Phycology, 50(3), 437-461.

70. Su, Y., Mennerich, A., Urban, B. 2011. Municipal wastewater treatment and biomass

accumulation with a wastewater-born and settleable algal-bacterial culture. Water Res,

45(11), 3351-3358.

71. Sun, J., Hu, Y.Y., Li, W.J., Zhang, Y.P., Chen, J., Deng, F. 2015. Sequential decolorization of

azo dye and mineralization of decolorization liquid coupled with bioelectricity generation

using a pH self-neutralized photobioelectrochemical system operated with polarity reversion.

Journal of Hazardous Materials, 289, 108-117.

72. Talbot, P., Thébault, J.-M., Dauta, A., De la Noüe, J. 1991. A comparative study and

mathematical modeling of temperature, light and growth of three microalgae potentially

useful for wastewater treatment. Water research, 25(4), 465-472.

73. Tanabe, Y., Kato, S., Matsuura, H., Watanabe, M.M. 2012. A Botryococcus Strain with

Bacterial Ectosymbionts Grows Fast and Produces High Amount of Hydrocarbons. Procedia

36
Environmental Sciences, 15, 22-26.

74. Tate, J.J., Gutierrez-Wing, M.T., Rusch, K.A., Benton, M.G. 2013. The Effects of Plant

Growth Substances and Mixed Cultures on Growth and Metabolite Production of Green

Algae Chlorella sp.: A Review. J Plant Growth Regul, 32(2), 417-428.

75. Ternes, T.A. 1998. Occurrence of drugs in German sewage treatment plants and rivers. Water

Research, 32(11), 3245-3260.

76. Ueshima, M., Ginn, B.R., Haack, E.A., Szymailowski, J.E.S., Fein, F.B. 2008. Cd adsorption

onto Pseudomonas putida in the presence and absence of extracellular polymeric substances.

Geochimica Et Cosmochimica Acta, 72(24), 5885-5895.

77. Valderrama, L.T., Del Campo, C.M., Rodriguez, C.M., de- Bashan, L.E., Bashan, Y. 2002.

Treatment of recalcitrant wastewater from ethanol and citric acid production using the

microalga Chlorella vulgaris and the macrophyte Lemna minuscula. Water Research, 36(17),

4185-4192.

78. Van Den Hende, S., Carre, E., Cocaud, E., Beelen, V., Boon, N., Vervaeren, H. 2014.

Treatment of industrial wastewaters by microalgal bacterial flocs in sequencing batch

reactors. Bioresour Technol, 161, 245-254.

79. Vanerkar, A.P., Fulke, A.B., Lokhande, S.K., Giripunje, M.D., Satyanarayan, S. 2015.

Recycling and treatment of herbal pharmaceutical wastewater using Scenedesmus

quadricuada. Current Science, 108(5), 979-983.

80. Vannini, C., Domingo, G., Marsoni, M., De Mattia, F., Labra, M., Castiglioni, S., Bracale, M.

2011. Effects of a complex mixture of therapeutic drugs on unicellular algae

Pseudokirchneriella subcapitata. Aquat Toxicol, 101(2), 459-465.

81. Vijayaraghavan, G., Shanthakumar, S. 2015. Removal of Sulphur Black Dye from its

Aqueous Solution Using Alginate from Sargassum sp (Brown Algae) as a Coagulant.

Environmental Progress & Sustainable Energy, 34(5), 1427-1434.

82. Vijayaraghavan, K., Yun, Y.S. 2008. Bacterial biosorbents and biosorption. Biotechnol Adv,

26(3), 266-291.

83. Wang, B., Lan, C.Q. 2011. Biomass production and nitrogen and phosphorus removal by the

green alga Neochloris oleoabundans in simulated wastewater and secondary municipal

wastewater effluent. Bioresour Technol, 102(10), 5639-5644.

37
84. Wang, H., Xiong, H., Hui, Z., Zeng, X. 2012. Mixotrophic cultivation of Chlorella

pyrenoidosa with diluted primary piggery wastewater to produce lipids. Bioresour Technol,

104, 215-220.

85. Wang, M., Yang, H., Ergas, S.J., van der Steen, P. 2015a. A novel shortcut nitrogen removal

process using an algal-bacterial consortium in a photo-sequencing batch reactor (PSBR).

Water Research, 87, 38-48.

86. Wang, Y., Guo, W., Yen, H.W., Ho, S.H., Lo, Y.C., Cheng, C.L., Ren, N., Chang, J.S. 2015b.

Cultivation of Chlorella vulgaris JSC-6 with swine wastewater for simultaneous

nutrient/COD removal and carbohydrate production. Bioresour Technol, 198, 619-625.

87. Wilkie, A.C., Mulbry, W.W. 2002. Recovery of dairy manure nutrients by benthic freshwater

algae. Bioresour Technol, 84(1), 81-91.

88. Wingender, J., Neu, T.R., Flemming, H.-C. 2012. Microbial extracellular polymeric

substances: characterization, structure and function. Springer Science & Business Media.

89. Wu, Y., Hu, Z., Yang, L., Graham, B., Kerr, P.G. 2011. The removal of nutrients from

non-point source wastewater by a hybrid bioreactor. Bioresour Technol, 102(3), 2419-2426.

90. Xiong, J.Q., Kurade, M.B., Abou-Shanab, R.A.I., Ji, M.K., Choi, J., Kim, J.O., Jeon, B.H.

2016. Biodegradation of carbamazepine using freshwater microalgae Chlamydomonas

mexicana and Scenedesmus obliquus and the determination of its metabolic fate.

Bioresource Technol, 205, 183-190.

91. Xu, M., Li, P., Tang, T., Hu, Z. 2015. Roles of SRT and HRT of an algal membrane

bioreactor system with a tanks-in-series configuration for secondary wastewater effluent

polishing. Ecological Engineering, 85, 257-264.

92. Xu, Y.J., Wang, Y., Yang, Y., Zhou, D.D. 2016. The role of starvation in biomass harvesting

and lipid accumulation: Co-culture of microalgae-bacteria in synthetic wastewater.

Environmental Progress & Sustainable Energy, 35(1), 103-109.

93. Yang, C., Hua, Q., Shimizu, K. 2000. Energetics and carbon metabolism during growth of

microalgal cells under photoautotrophic, mixotrophic and cyclic

light-autotrophic/dark-heterotrophic conditions. Biochem Eng J, 6(2), 87-102.

94. Zhang, W., Zhang, M., Lin, K., Sun, W., Xiong, B., Guo, M., Cui, X., Fu, R. 2012.

Eco-toxicological effect of carbamazepine on Scenedesmus obliquus and Chlorella

38
pyrenoidosa. Environ Toxicol Pharmacol, 33(2), 344-352.

95. Zhao, S., Gao, B.Y., Yue, Q.Y., Wang, Y., Li, Q., Dong, H.Y., Yan, H. 2014. Study of

Enteromorpha polysaccharides as a new-style coagulant aid in dye wastewater treatment.

Carbohydrate Polymers, 103, 179-186.

96. Zhou, D., Niu, S., Xiong, Y., Yang, Y., Dong, S. 2014. Microbial selection pressure is not a

prerequisite for granulation: dynamic granulation and microbial community study in a

complete mixing bioreactor. Bioresour Technol, 161, 102-108.

97. Zhou, W., Li, Y., Min, M., Hu, B., Chen, P., Ruan, R. 2011. Local bioprospecting for

high-lipid producing microalgal strains to be grown on concentrated municipal wastewater

for biofuel production. Bioresour Technol, 102(13), 6909-6919.

39
Table 1 Symbiotic association between algae and bacteria in the algae-bacteria consortium

Benefits Drawbacks
• CO2 from bacterial • Algicidal effects of some
metabolism bacteria
• Stimulative effects and
Algae
essential nutrients
from bacterial
metabolism
• Enhanced flocculation
by associated bacteria
• Oxygenation from • Increase in pH due to
algae associated algal
metabolism
Bacteria • Algal Organic matter • Increase in temperature
as a carbon source due to associated algal
metabolism
• Antibacterial effects
from some algae

40
Table 2 Different microalgae-bacteria consortia used in wastewater treatment

Pollutant removal
Microalgae-bacteria consortium Wastewater COD Nitrogen Phosphorus Reference
Note
(%) (%) (%)

Chlorella vulgaris + activated sludge Synthetic wastewater 83.6 89.4 91.4 Settling ability increased compared (Xu et al.,

with pure microalgae 2016)

Consortium of algae, consisting primarily of Anaerobically digested NA NA 90 Photo-sequencing batch reactor (Wang et

Chlorella (95.2%), Chlamydomonas (3.1%), swine manure (PSBR); using organic carbon source al., 2015a)

and Stichococcus (1.1%) + bacteria

Chlorella vulgaris+ Azospirillum brasilense Synthetic wastewater NA 99 83 (de-Basha

n, 2002)

Chlorella sorokiniana+Azospirillum brasilense Municipal wastewater NA NA 72 (Hernande

z et al.,

2006)

Chlorella. sorokiniana+Azospirillum brasilense Ammonia wastewater NA 100 NA Higher temperatures (40oC) and (de-Basha

intensity of light (2500 µmol m-2 s-1) n et al.,

2008)

Chlorella vulgaris+ Lemna minuscula Recalcitrant effluent 61 71.6 28 Ethanol and citric acid production (Valderra

41
ma et al.,

2002)

Scenedesmus sp. + Bacteria group Municipal wastewater 92.3 95.7 98.1 Bacteria: Flavobacteria and (Lee et al.,

Sphingobacteria 2016)

Chlorella. sorokiniana+ aerobic sludge Swine wastewaters 62.3 82.7 58.0 Nitrification efficiency: 75.7 % (Hernande

denitrification efficiency: 53.8% z et al.,

2013)

Oscillatoria sp. OSC+Proteobacteria naturally Oil compounds in NA NA NA n-Octadecane 40% (Abed &

associated with Oscillatoria sp. wastewater pristine 50% Koster,


phenanthrene 50% 2005)
dibenzothiophene 80%

Cyanobacterial mats+Acinetobacter Oil compounds NA NA NA Oil 63.2% (0.5% v/v) (Al-Awad

calcoaceticus and Nocardioforms hi et al.,

2003)
Sodium salicylate removal
Chlorella sorokiniana 211/8k+Ralstonia Toxic compounds NA (Guieysse
1 mmol l-1 day
basilensis et al.,

2002)

42
Table 3 Microalgae used in the treatment of pharmaceutical wastewaters

Species Compounds Biodegradation Note Reference

Chlamydomonas carbamazepine 35% 97% growth inhibition at (Xiong et


-1
mexicana 100 mg L al., 2016)

Scenedesmus obliquus carbamazepine 28% 30% growth inhibition at (Xiong et


100 mg L-1 al., 2016)

Chlamydomonas carbamazepine NA 31.3% growth inhibition (Xiong et


pitschmannii at 100 mg L-1 al., 2016)

Miractinium resseri carbamazepine NA 43.5% growth inhibition (Xiong et


-1
at 100 mg L al., 2016)

Scenedesmus obliquus carbamazepine NA EC50 was 54.60 mg L-1 (Zhang et


al., 2012)

Chlorella pyrenoidosa carbamazepine NA EC50 was 33.11 mg L-1 (Zhang et


al., 2012)

Scenedesmus Pharmaceutical Absorption Microalgae can tolerance (Vanerkar


a
quadricauda wastewater 20% PCTE et al., 2015)

Chlorella kessleri Pharmaceutical NA 13 psychoactive (Mackulak


wastewater pharmaceuticals were et al., 2015)
selected for experiment.

Microalgae consortium Wastewater containing up to 90% 4-octylphenol, galaxolide, (Matamoros


emerging contaminants and tributyl phosphate et al., 2016)
concentrations

Microalgae consortium Wastewater containing 17% caffeine (Matamoros


emerging contaminants et al., 2016)

Consortia of microalgae Wastewater containing 99% caffeine (Matamoros

43
and bacteria in emerging contaminants et al., 2016)
wastewater

Microalgae consortium Wastewater containing 15% ibuprofen (Matamoros


emerging contaminants et al., 2016)

Consortia of microalgae Wastewater containing 60% ibuprofen (Matamoros


and bacteria in emerging contaminants et al., 2016)
wastewater

Consortia of microalgae Wastewater containing < 20% Carbamazepine and (Matamoros


and bacteria in emerging contaminants tris(2-chloroethyl) et al., 2016)
wastewater phosphate

a: PCTE: Physico-chemically treated effluent.

NA: not available.

44
Table 4 Microalgae used for the treatment of textile dye from wastewaters

Species Compounds mechanism Note Reference


Chlorella vulgaris (immobilized in Lanaset Red 2GA Adsorption 44% dye removal from (Chu et al., 2009)
2% alginate) initial concentration of 7.5
mg L-1

Chlorella vulgaris Supranol Red 3BW Adsorption High rate algae ponds (Zhou et al., 2014)
50% color removal

Chlorella vulgaris mono-azo dye yellow 2G Bio-sorption 63-90% color removal (Acuner & Dilek, 2004)

Chlorella vulgaris Remazol Black B, Remazol Red Adsorption (Aksu & Tezer, 2005)
RR, Remazol Golden Yellow

Chlorella sp. Indigo textile dye Adsorption 46% color removal (Cheriaa et al., 2009)

Caulerpa lentillifera Astrazon Blue FRGL Adsorption Dried algae biomass used (Marungrueng & Pavasant, 2006)
as biosorbent

Caulerpa lentillifera CI Basic Blue, CI Basic Red, CI Adsorption Algae biomass can (Marungrueng & Pavasant, 2007)
Basic Blue sequester Red GTLN more
rapidly when compared
with activated carbon

45
Synechocystis and Phormidium reactive dye Bio-adsorption Cyanobacteria can be used (Karacakaya et al., 2009)
for dye removal with
stimulation of biomass
production

Phormidium.sp Remazol Blue and Reactive Adsorption 88% color removal (Ertugrul et al., 2008)
Black B
Spirogyra Synazol Adsorption 85% decolorizaiton with (Khalaf, 2008)
dried biomass

Hypnea valentiae dye Adsorption 88% decolorizaiton with (Khalaf, 2008)


dried biomass
Nostoc linckia Methyl red Degradation 82% color removal (El-Sheekh et al., 2009)

Lyngbya lagerlerimi Orange II Degradation 47% color removal (El-Sheekh et al., 2009)

Nostoc linckia Basic cationic Degradation 92% color removal (El-Sheekh et al., 2009)

Chlorella vulgaris G-Red (FN-3G) Degradation 59% color removal (El-Sheekh et al., 2009)

Oscillatoria rubescens Basic Fuschin Degradation 95% color removal (El-Sheekh et al., 2009)

46
Table 5 Photobioreactors (PBR) used for wastewater treatment using microalgae-bacteria consortia

PBR Type Strain Nitrogen Note Reference


removal

Tubular biofilm S. quadricauda+ enriched 0.092 g/L/day 1 g of N input supports the production of 16.8 g of (Gonzalez et al., 2008)
culture of nitrifiers oxygen, which is sufficient to oxidise another 5.6 g
of nitrogen.

Tubular biofilm Microalgae+bacteria Below 0.01 mg Dominate microalgae: Chlorella, Oocystis and (Krustok et al., 2015)
-1
L in effluent Scenedesmus. Dominate bacteria: Aulacoseira,
Stephanodiscus, Diatoma, Cryptophyceae and
Melosira

Photo-sequencing Chlorella + bacteria 90% Without aeration (Wang et al., 2015a)


batch reactor
(PSBR)

Sequencing batch consortium of NA Microalgal bacterial flocs; (Van Den Hende et al., 2014)
reactors (MaB-floc microalgae/cyanobacteria
Four industrial wastewater
SBRs)

High rate algae microalgae-bacteria 80% slaughterhouse wastewater, working volume 75 L


ponds consortia

47
CO2
O2
(a) H2, CH4, CO2

DO

Wastewater
Effluent
Anaerobic Aerobic
treatment treatment
Biomass
Organic acid, CO2
Biodisel

Fermentation

Digestion

Fertilizer

Fodder

(b)

Fig. 1 COD/nutrient remvoal using of microalgae-bacteria system, (a) integrated

processes, (b) nutrient and energy flow in microalgae-bacteria consortium

48
vlgaris
Cyanobacteria sorokiniana Municipal wastewater
Chlorella Fish farm wastewater
pyrenoidosa
Chlamydomonas Artificial wastewater
minutissima slaughterhouse wastewater
Cosmarium
kessleri Anaerobic digestion effluent
Desmodesmus livestock wastewater
Chlorophyta protothecoide
Ankistrodesmus Textile dyes wastewater
zofingiensis Swine industry wastewater
obliquus
Scenedesmus dimorphus Pharmaceutical wastewater dairy wastewater
acutus
Dunaliella
platensis Agro-industrial wastewater soybean processing wastewater
quadricauda
Neochloris electroplating wastewater
oleoabundans
Metal containing wastewater
pseudoalveolaris coal-fired waste water
Bacillariophyta Nitzschia
Industrial effluent(ethanol
Euglenophyta Euglena and citric acid production)
Lemna minuscula
Macroalgae
Sargassum Sugar mill effluent
Pelvetia canaliculata
Ascophyllum
Fucus spiralis
Laminaria hyperborea
Oedogonium

Fig. 2 Microalgal species used in various wastewater treatment

49
Textile dye wastewater
Heavy metal Biodegradation

Dye compound
Adsorption
precipitation
Bioadsorption for
+ + decolorizaiton
+

Accumulation
CO2
CO2 Redox enzyme

CO2 Bacteria EPS increasing

Aerobic bacteria O Intermediates with low


2 toxicity
O2
Organic Antioxidative
defense
carbon mechanisms
N source
P source
Agro-industrial wastewater Pharmaceutical wastewater

Fig. 3 Mechanisms of various industrial wastewater treatments using microalgae and bacteria

50
Highlights

• Four industrial wastewaters treated by microalgae-bacteria consortia were

reviewed

• Photobioreactor design for wastewater treatment with microalgae were

described

• Feasibility and potential of microalgae-based wastewater treatment was

evaluated

You might also like