You are on page 1of 11

Bioresource Technology 303 (2020) 122886

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Review

Bioremediation of heavy metals using microalgae: Recent advances and T


mechanisms

Yoong Kit Leonga, Jo-Shu Changa,b,c,
a
Department of Chemical and Materials Engineering, College of Engineering, Tunghai University, Taichung, Taiwan
b
Department of Chemical Engineering, National Cheng Kung University, Tainan, Taiwan
c
Center for Nanotechnology, Tunghai University, Taichung, Taiwan

A R T I C LE I N FO A B S T R A C T

Keywords: Five heavy metals namely, arsenic (As), cadmium (Cd), chromium (Cr), lead (Pb) and mercury (Hg) are carci-
Microalgae nogenic and show toxicity even at trace amounts, posing threats to environmental ecology and human health.
Bioremediation There is an emerging trend of employing microalgae in phycoremediation of heavy metals, due to several
Biosorption benefits including abundant availability, inexpensive, excellent metal removal efficiency and eco-friendly
Heavy metals
nature. This review presents the recent advances and mechanisms involved in bioremediation and biosorption of
Mechanism
these toxic heavy metals utilizing microalgae. Tolerance and response of different microalgae strains to heavy
metals and their bioaccumulation capability with value-added by-products formation as well as utilization of
non-living biomass as biosorbents are discussed. Furthermore, challenges and future prospects in bioremediation
of heavy metals by microalgae are also explored. This review aims to provide useful insights to help future
development of efficient and commercially viable technology for microalgae-based heavy metal bioremediation.

1. Introduction 2019). Hence, it is crucial to remediate these toxic heavy metals in


wastewater effluents before discharging them.
“Heavy metals” is a term generally referred to metals and metalloids Conventional techniques employed for heavy metals removal from
with density more than 5 g/cm3 which include arsenic (As), cadmium wastewater include chemical precipitation, ion exchange, membrane
(Cd), chromium (Cr), copper (Cu), iron (Fe), lead (Pb), mercury (Hg), filtration, floatation, coagulation-flocculation and electrochemical
silver (Ag), zinc (Zn) and others (Duruibe et al., 2007). They are widely methods (such as electrodeposition, electrofloatation and electro-
involved in human activities, such as fossil fuel combustion, mining, coagulation) (Fu & Wang, 2011). However, these techniques have the
electroplating, dye and pigments manufacturing, fertilizers, and other disadvantages of expensive operation and maintenance costs, as well as
industrial activities, which are then released in large amount into the secondary pollution due to toxic sludge formation. Furthermore, most
environment daily via wastewater or other pathways (Zakhama et al., of these techniques are further expensive and/or ineffective at very low
2011). Due to their non-biodegradable characteristics, heavy metals concentrations of heavy metals, specifically below 100 mg/L (Birungi &
tend to persist in nature, leading to bio-accumulation in food chain Chirwa, 2015; Jaafari & Yaghmaeian, 2019).
which causes severe environmental and health issues (Yang et al., On account of their environmental-friendliness and cost-efficiency
2015). at low heavy metals concentration, bioremediation of heavy metals
Following the regulations of US Environmental Protection Agency, using different microorganisms including bacteria, microalgae, yeasts
the maximum contamination limits (MCL) of As, Cd, Cr, Hg and Pb are and fungi have been widely applied as alternatives to conventional
0.01, 0.005, 0.1, 0.002 and 0.015 mg/L respectively (USEPA, 2016). methods. Among all microorganisms, microalgae which has out-
These heavy metals and their compounds, even at very low con- standing biological characteristics such as high photosynthetic effi-
centrations, are highly toxic, carcinogenic, mutagenic and teratogenic. ciency and simple structure has the ability to grow well under extreme
Direct contact, inhalation and ingestion of these heavy metals poses environmental conditions such as presence of heavy metals, high sali-
serious threats to human physical and mental health, causing mutations nity, nutrient stress and extreme temperature. Therefore, there is an
and genetic damage, damaging central nervous system as well as es- emerging trend of employing microalgae in phycoremediation of toxic
calating the risk of cancers (Balaji et al., 2016a; Jaafari & Yaghmaeian, heavy metals due to its high binding affinity, abundance of binding sites


Corresponding author at: Department of Chemical and Materials Engineering, College of Engineering, Tunghai University, Taichung, Taiwan.
E-mail address: changjs@mail.ncku.edu.tw (J.-S. Chang).

https://doi.org/10.1016/j.biortech.2020.122886
Received 20 December 2019; Received in revised form 20 January 2020; Accepted 21 January 2020
Available online 23 January 2020
0960-8524/ © 2020 Elsevier Ltd. All rights reserved.
Y.K. Leong and J.-S. Chang Bioresource Technology 303 (2020) 122886

and large surface area (Cameron et al., 2018). Furthermore, both living activity and ascorbic acid-glutathione (ASC-GSH) pathway as well as
cells and non-living biomass of microalgae can be employed as bio- sustaining the dissipation of surplus excitation energy and scavenging
sorbents. Other than outstanding removal capacity and being eco- ROS. Furthermore, high level of ASC is secreted by microalgae as a
friendly, bioremediation of heavy metals using microalgae has the hydrophilic redox buffer that is responsible for the protection of cytosol
benefits of robust and simple process, lack of toxicity constraint, rapid and other cellular components against oxidative threats. On the other
growth rate compared to higher plants as well as formation of value- hands, high level of GSH protect the microalgae by providing tolerance,
added products such as biofuels and fertilizers (Abinandan et al., 2019; scavenging free radicals, facilitating PCs and ASC synthesis as well as
Balaji et al., 2016b). In addition, oxidative stress induced by heavy restoration of substrate for other antioxidants (Gómez-Jacinto et al.,
metals enhances the lipid content of microalgae (Upadhyay et al., 2015; Upadhyay et al., 2016).
2016). Furthermore, microalgae can also be employed for the recovery Removal of heavy metals by microalgae is attained via a two-stage
of precious metal ions, such as thallium, silver and gold (Birungi & mechanism. The first stage is the rapid extracellular passive adsorption
Chirwa, 2015; Jaafari & Yaghmaeian, 2019). (biosorption), while the second stage is the slow intracellular positive
In this review, we present the recent advances and in-depth sum- diffusion and accumulation (bioaccumulation). In addition to cell
mary of the development of bioremediation of five non-threshold toxic polymeric substances such as peptides and exopolysaccharides with
heavy metals, which are arsenic (As), cadmium (Cd), chromium (Cr), uronic groups, cell wall of microalgae are mainly composed of poly-
lead (Pb) and mercury (Hg) by microalgae. These heavy metals have saccharides (cellulose and alginate), lipids and organic proteins, pro-
gained much attention and became a significant research focus due to vides many functional groups (such as amino, carboxyl, hydroxyl,
their severe deteriorating effects to the environmental ecology and imidazole, phosphate, sulfonate, thiol and others) capable of binding
human health (Rahman & Singh, 2019; Suvarapu & Baek, 2017). heavy metals (Priatni et al., 2017). Furthermore, they consisted of huge
amount of monomeric alcohols, laminaran, deprotonated sulphate and
2. Mechanisms of heavy metal removal by microalgae carboxyl groups which attract both anionic and cationic species of
different heavy metals. Microalgae biomass also has different functional
Microalgae consume heavy metals such as boron (B), cobalt (Co), groups, such as amide, carbonyl, carboxylic acid, ether and hydroxyls
copper (Cu), iron (Fe), molybdenum (Mo), manganese (Mn), and zinc which contribute to the biosorption (Pradhan et al., 2019). Fourier-
(Zn) as trace elements for enzymatic process and cell metabolism, transform infrared (FTIR) analysis and nuclear magnetic resonance
though other heavy metals such as As, Cd, Cr, Pb and Hg are toxic to (NMR) spectroscopy can be employed to determine the functional
microalgae. Due to hormesis phenomenon, low toxic heavy metal groups in microalgae that are responsible for formation of complexes
concentrations can stimulate the growth and metabolism of microalgae with heavy metals as well as examine the interaction between them.
(Sun et al., 2015). Some cyanobacterial species, such as Anabaena, The knowledge of biosorbents chemical structure provides important
Oscillatoria, Phormidium and Spirogyra can naturally grow in water information for prediction of their affinities for heavy metal ions.
contaminated by heavy metals due to their tolerance towards heavy Adsorption of heavy metal on the surface of microalgae is a rapid
metal stress (Balaji et al., 2016a). Other than heavy metals, microalgae process and can happen via different paths, which are formation of
have reactive groups with active binding sites which can form com- covalent bond between ionized cell wall with heavy metals, ionic ex-
plexes with pollutants in wastewater. This leads to flocculation and change of heavy metal ions with cell wall cation and binding of heavy
subsequently reduces the content of both total dissolved solids (TDS) metal cations with negatively charged uronic acids of microalgae exo-
and total suspended solids (TSS) (Balaji et al., 2016a). polysaccharides. On the other hand, the process of heavy metal accu-
Microalgae species has multiple strategies of unique self-protection mulation inside the cell is much slower. The heavy metals are actively
mechanisms against toxicity of heavy metals, such as heavy metal im- transported across the cell membrane and into the cytoplasm followed
mobilization, gene regulation, exclusion and chelation as well as anti- by diffusion and subsequent binding with internal binding sites of
oxidants or reducing enzymes which reduce heavy metals via redox proteins and peptides such as GSH, metal transporter, oxidative stress
reactions (Gómez-Jacinto et al., 2015). Microalgae can form cellular reducing agents and phytochelatins (Ibuot et al., 2017; Pradhan et al.,
protein-heavy metals complexes without changing its own activity 2019). Fig. 1 illustrates the mechanisms of heavy metals removal by
(Priatni et al., 2017). The organometallic complexes are further sepa- microalgae.
rated insides vacuoles to help regulate the heavy metal ions con-
centration in cytoplasm which subsequently mitigate their toxic effects 3. Bioremediation and biosorption of heavy metals
(Balaji et al., 2016b). In addition, heavy metals activate the biosynth-
esis of phytochelatins (PCs) which are thiol-rich peptides and proteins 3.1. Arsenic (As)
that minimize heavy metal stress by interacting with them (Gómez-
Jacinto et al., 2015). Classified by the United States Environmental Protection Agency
To counteract with free radicals released by heavy metals during (USEPA) as class A and category 1 carcinogen, arsenic (As) is one of the
adsorption, microalgae synthesize antioxidant enzymes such as ascor- major toxic heavy metals that cause contamination of potable water in
bate peroxidase, catalase, glutathione reductase, peroxidase and su- many countries, such as Argentina, Bangladesh, Chile, China, Finland,
peroxide dismutase (SOD) as well as non-enzymatic antioxidants such India, South East Asia and USA (Arora et al., 2018; Upadhyay et al.,
as carotenoids, cysteine, ascorbic acid (ASC), glutathione (GSH) and 2016; Wang et al., 2014). Widespread pollution of arsenic is caused by
proline (Upadhyay et al., 2016). SOD acts as the first line of defence anthropogenic activities, such as combustion of fossil fuels, mining,
against superoxide anion by breaking it down into oxygen molecules medical use, fertilizers and pesticides, smelting, electrolytic process,
and hydrogen peroxide. The hydrogen peroxide is further degraded by sewage sludge as well as pigments, semiconductor, glass and alloy
catalase into water and oxygen molecules (Balaji et al., 2016b). Cy- manufacturing processes (Arora et al., 2017; Singh et al., 2016). Phy-
steine indirectly or directly serves as the precursor for PCs, GSH, me- sical exposure to As even at low concentrations (0.1 µg/mL) cause lung,
tallothioneins and other sulphur containing compounds, hence serving skin, kidney and bladder cancers, while higher concentration causes
as the indicator for the synthesis of different antioxidants. arsenicosis, arsenical dermatitis, cardiovascular disease, diabetes, in-
GSH and ASC are important endogenous antioxidants synthesized fant morbidity, immune alteration, impairment to the central nervous
by microalgae and they play a key role in the reduction of reactive system, hepatic damage and hyperkerotosis (Li et al., 2019). Con-
oxygen species (ROS) and free radicals (Devars et al., 2000). Other than versely, in plant and microalgae, arsenic causes impairment to cell or-
maintaining the equilibrium of ROS production and elimination, ASC ganelle and DNA, lipid peroxidation and protein degradation
protects microalgal cells by regulating metal-containing enzyme (Upadhyay et al., 2016).

2
Y.K. Leong and J.-S. Chang Bioresource Technology 303 (2020) 122886

Fig. 1. Mechanism of heavy metals removal by microalgae.

The toxicity and physico-chemical properties of arsenic is depen- fungal-algal pellets and high removal capacity was achieved, though
dent on it oxidation state and chemical form. Arsenic has four oxidation arsenite methylation did not occur (Li et al., 2019). Wang and co-
states, which are arsenic (As0), arsine [As(-III)], arsenite [As(III)] and workers reported that PO43− and As(V) has a synergistic interaction
arsenate [As(V)]. The most commonly occurring oxidation states of As and significantly promoted the initial growth of axenic cultures of Du-
are trivalent [As (III)] and pentavalent [As(V)] forms, of which the naliella salina, though the growth experienced inhibition after 9 days of
former is more toxic (Zhu et al., 2018). In aqueous solution, As (III) cultivation. The presence of symbiotic bacteria and PO43− limitation
exists predominantly as H3AsO3 with other forms such as H4AsO3+, are shown to promote the reduction of As(V) and induce the excretion
H2AsO3−, HAsO32− and AsO33−, while As(V) exists as H3AsO4, of As (III) and DMA by D. salina (Wang et al., 2016).
H2AsO4−, HAsO42− and AsO43− (Arora et al., 2017). The inorganic Different freshwater and marine microalgae have varying tolerance
form of arsenic is much more toxic than its organic form. Thus, it is to arsenic, thus experience different growth rate and bioaccumulation
beneficial to transform arsenic into its organic form, which includes capacity in the presence of arsenic. Examining the bioaccumulation of
monomethylarsonate (MMA), dimethylarsinate (DMA), arsenocholine arsenate by marine microalgae Phaeoductylum tricornutum, Morelli and
(ArsC) and arsenobetaine (ArsB) (Wongrod et al., 2019). Microalgae colleagues observed a gradual decrease in growth rate when the con-
reduces the toxicity of inorganic arsenic by As (III) oxidation and centration of As(V) was above 0.1 µm and the microalgae experienced
complex formation with phytochelatins and glutathione, As(V) reduc- maximum inhibition of 35% at 1.0 µm arsenic (Morelli et al., 2005).
tion, biotransformation into arsenolipids/arsenosugars or methylated Comparing between two freshwater microalgae, Levy and co-workers
arsenic species, extracellular adsorption and excretion from the cell found that Chlorella sp. has more tolerance to arsenic compared to
(Papry et al., 2019). Oxidation of As(III) mainly happens outside the Monoraphidium arcuatum (Levy et al., 2005). Jiang and co-workers re-
cells by interacting with functional groups such as –OH, –NH, –CH of ported high tolerance of Chlorella vulgaris to As(V) up to 200 mg/L with
aldehyde and aliphatic as well as –CN of amino groups (Arora et al., constant arsenic removal of 70%. Extracellular heavy metal adsorption
2018; Wang et al., 2014). On the other hand, cellular metabolism of serves a critical role in As(V) removal by microalgae at concentrations
arsenic starts with rapid reduction of As(V) to As (III) followed by a below 100 mg/L (Jiang et al., 2011). Huang and colleagues observed
slower step-by-step methylation first to MMAs, and then to DMAs. cyanobacterial species such as Anabaena affinis, Microcystis aeruginosa
Microalgae have been shown to excrete both methylated arsenic species and Oscillatoria tenuis experienced up to 2.6-fold slower growth rate in
and/or reduced As(III) (Baker & Wallschläger, 2016). As(V)-containing medium compared to control. Among these micro-
Essential for protein synthesis, regulation of genetic materials and algae, A. affinis and O. tenuis removed As(V) by binding them to ex-
modification of protein via phosphorylation, phosphate (PO43−) plays tracellular cell wall active sites, while intracellular bioaccumulation in
an important role in speciation, bioaccumulation and detoxification of the cytoplasm for most arsenic removal occurred in M. aeruginosa
arsenic in microalgae due to the similarity in outer electron config- (Huang et al., 2014). Upadhyay and colleagues has reported that
uration as arsenate [As(V)] (Sun et al., 2015). Therefore, the high af- Nannochloropsis sp. can grow well in As (III) concentration up to 100 µm
finity phosphate transporters of microalgae (such as PIT and PST) can and have a significant decrease in biomass content at As (III) con-
participate in the uptake of arsenic. PO43− limitation induces the centration further increases to 500 µm. Rapid decrease in biomass at
synthesis of arsenic transporter and thus, accelerates arsenic uptake high concentration of arsenic might be caused by the blocking of -SH
(Wang et al., 2014). On the other hand, increased phosphate con- group, leading to disruption of metabolism by disintegration of essen-
centration significantly reduced the arsenic uptake in microalgae such tial cellular proteins (Upadhyay et al., 2016). Singh and colleagues also
as Chlorella sp. (Bahar et al., 2016), C. vulgaris (Baker & Wallschläger, reported high accumulation of arsenic up to 760 µg/g in microalgae
2016; Li et al., 2019), C. salina (Karadjova et al., 2008), Dunaliella salina species such as Hydrodictiyon reticulatom, Diatoms, Pithophora sp.,
(Wang et al., 2017; Wang et al., 2016), Chlamydomonas reinhardtii and Phormidium sp. and Oscillatoria sp. found in As-contaminated area,
Scenedesmus obliquus (Wang et al., 2013). This is because both As(III) showing potential in bioremediation of As (Singh et al., 2016). Arora
and As(V) experienced non-competitive and competitive inhibition re- and co-workers found that Chlorella minutissima and Scenedesmus sp.
spectively, by extracellular PO43−. However, it is reported that PO43− IITRIND2 can tolerate up to 500 mg/L arsenic with removal up to
concentration has negligible impact on the removal of arsenic by 161 µg/g (Arora et al., 2017). Comparing between four green

3
Y.K. Leong and J.-S. Chang Bioresource Technology 303 (2020) 122886

microalgae species (Chlorella vulgaris, Chlamydomonas reinhardtii, Sce- 249.36 mg/L day was also higher than many reported in literature
nedesmus almeriensis and an indigenous Chlorophyceae spp.), S almer- (Yang et al., 2015). Monoraphidium sp. QLY-1 treated with 80 µm Cd(II)
iensis demonstrated highest removal of As up to 41.7% at pH 9.5 showed enhanced lipid content (52.78%) of 1.59 fold compared to
(Saavedra et al., 2018). control (Zhao et al., 2019). A 0.02 µm of free Cd(II) improved growth
In addition, microalgae respond differently under oxidative stress rate and total lipids of Chlorella vulgaris by 13.4% and 51.5% respec-
induced by arsenic. Responding to arsenic stress, decrease in protein tively (Chia et al., 2013). Similarly, acid-tolerant microalgae, Hetero-
and chlorophyll content was observed in both C. minutissima and chlorella sp. MAS3 and Desmodesmus sp. MAS1 can simultaneously re-
Scenedesmus sp. In addition, cell size reduction was recorded in C. move 2 mg/L Cd(II) with > 58% efficiency and accumulated lipids rich
minutissima, while Scenedesmus sp. has increased lipid content at the in aliphatic long-chain hydrocarbons and triglycerides at pH 3.5
cost of carbohydrate reserve (Arora et al., 2017). Upon bioaccumula- (Abinandan et al., 2019).
tion of arsenic, the total chlorophyll content of Nannochloropsis sp. With the advantage of cost-effective biomass harvesting from di-
decreased, while carotenoids and lipid content increased. Nanno- luted culture, self-flocculating microalgae, such as Chlorella vulgaris
chloropsis sp. treated with 100 µm As(III) has the highest lipid pro- JSC-7 (Alam et al., 2015) and Scenedesmus obliquus AS-6-1 (Zhang et al.,
ductivity of 20.27 mg/L/day and produce biofuel meeting EN 14214 2016), have been utilized for the removal of Cd(II). Due to the dis-
standards (Upadhyay et al., 2016). Sun and co-workers observed an tinctive cell wall components, enhanced synthesis of phytohormones
increase in the external cell surface area of Nannochloropsis sp. due to and higher antioxidation activity, C. vulgaris JSC-7 was more tolerant to
formation of measles-like granules and wrinkles caused by arsenic Cd2+ with an improved heavy metal removal capacity of 21.5 mg/g
toxicity. The presence of arsenic increased the accumulation of short- compared to 6.5 mg/g of non-flocculating control species, (i.e., C. vul-
chain and monounsaturated fatty acids which is desirable for high garis CNW11) (Alam et al., 2015). Other than the main contribution
quality biofuel production (Sun et al., 2015). Scenedesmus sp. IITRIND2 from phosphate and carboxyl groups, functional groups such as amide,
cultured in As-containing synthetic soft water produced microalgal lipid C-N, hydroxyl and S-O groups also responsible for adsorption of Cd(II)
with linoleic acid, oleic acid, palmitic acid and stearic acid as major in S. obliquus AS-6-1. Successive adsorption/desorption cycle studies
constituents. The biodiesel produced from this lipid not only met the showed that S. obliquus AS-6-1 maintained its flocculating properties as
requirement of biodiesel standards (EN14214 and ASTM D6751-52), cadmium chloride did not impair the flocculating properties and might
but also has higher oxidative stability, leading to a longer shelf life even help induce chemical flocculation (Zhang et al., 2016).
(Arora et al., 2017). Table 1 summarizes the biosorption of arsenic (As) Different strategies such as genetic engineering, bio-immobilization,
by different microalgae strains. microalgae immobilization and bio-pellets have been studied to im-
proved cadmium biosorption. Gene manipulation has been performed
to overexpress metal tolerance protein in Chlamydomonas reinhardtii,
3.2. Cadmium (Cd) which showed a 2.29 and 3.06 fold increased in Cd2+ tolerance and
uptake, respectively, when compared with wild type C. reinhardtii
Characterized as a highly toxic invasive heavy metal, cadmium is (Ibuot et al., 2017). However, the natural strains adapted to wastewater
released into the environment through anthropogenic activities such as growth such as Parachlorella hussii, Parachlorella kessleri and Chlorella
waste incineration, manufacture of metals and alloys, ceramics, colour luteoviridis still demonstrated better heavy metal tolerance and accu-
pigment, fertilizers and pesticides, nickel-cadmium batteries, plastic, mulation. A novel technique for Cd ions bio-immobilization utilizing
and also electroplating, mining, smelting and welding process Microcystis aeruginosa and Selenastrum capricornutum cells as bio-tem-
(Abinandan et al., 2019). Due to its carcinogenic and teratogenic plates by co-incubating with selenium for bio-fabrication of fluorescent
nature, cadmium induces severe damage to reproductive organs, CdSe NPs has been performed. The reducing biomolecules of the mi-
kidney, liver and lungs even at trace amount, leading to health pro- croalgae, such as glutathione, NADPH and NADPH dependent reductase
blems such as cancer, Parkinson’s and Alzheimer’s diseases, gastro- were speculated to play a crucial role in the formation of CdSe NPs
intestinal disorder, amyotrophic lateral sclerosis, hypertension, kidney (Zhang et al., 2019). Immobilized Chlorella sp. in alginate bead has
damage, kidney stone, peripheral neuropathy, osteoporosis and re- demonstrated biosorption of Cd(II) up to 60% at low Cd2+ concentra-
spiratory insufficiency (Zhang et al., 2019). tion of 20 ppm (Valdez et al., 2018). In addition, bioremediation of
Microalgae strains respond differently under cadmium stress. In cadmium by Chlorella sp. – water hyacinth biochar complex achieved
addition to reducing cell growth and chlorophyll content, Cd(II) in- 214.7 mg/g higher than both Chlorella sp. and biochar alone due to
duces the production of phytochelatins, superoxide dismutase, catalase enhancement of the surface potential. Functional groups such as eOH,
and peroxidase as defence mechanism in microalgae such as eNH, P]O (phosphoryl group), P]S (phosphorotioate group), C]O of
Chlamydomonas moewusii and Monoraphidium sp. (Mera et al., 2016; amide and CeO of alcoholic groups contributes to cadmium biosorption
Zhao et al., 2019). On the other hand, Chlorella minutissima UTEX2341 (Shen et al., 2017). Similarly, Chlorella sp. immobilized with water-
demonstrated strong adaptability to cadmium with maximum adsorp- hyacinth leaf biochar pellets achieved 92.5% Cd(II) removal efficiency
tion capacity of 35.65 mg/g. Uptake of 0.6 mM Cd(II) have shown to and remained viable at 10 mg/L Cd(II) (Shen et al., 2018). Bio-pellets
enhance both microalgal lipid content and productivity up to 42.1% composed of C. vulgaris microalgae and Apsergillus niger fungi showed
and 2.17 fold compared to control. The maximum lipid productivity of

Table 1
Performance of biosorption of arsenic (As) by different microalgae strains.
Microalgae strains Temp (°C) Optimal pH Initial metal Biomass conc. Time Max. sorption Removal Reference
conc. (mg/L) (g/L) (min) (mg/g) efficiency (%)

Maugeotia genuflexa 20 6 10 4 60 2.4 96 (Sarı et al., 2011)


Mixture of green (Chlorophyta) algae and 20 4 50 10 180 3.5 70 (Sulaymon et al., 2013a;
blue-green (Cyanobacteria) algae Sulaymon et al., 2013b)
Spirulina sp. 35 6 7276 12 240 365 60.2 (Doshi et al., 2009)
Ulothrix cylindricum 20 6 10 4 60 2.45 98 (Tuzen et al., 2009)
Chlamydomonas reinhardtii – 9.5 12 1 180 4.63 38.6 (Saavedra et al., 2018)
Chlorella vulgaris – 5.5 12 1 180 3.89 32.4
Scenedesmus almeriensis – 9.5 12 1 180 5.0 41.7

4
Y.K. Leong and J.-S. Chang Bioresource Technology 303 (2020) 122886

better performance for the removal of Cd(II) at low level of 1 µg/L up to Several mechanisms have been proposed for the reduction and re-
56% compared to 40% of microalgae alone, with the advantages of moval of Cr(VI) by microalgae, depending on the binding properties of
lower final pH values and simpler harvesting (Bodin et al., 2017). the functional groups and the nature of operating conditions.
On the other hand, utilization of non-living microalgae cells for the Biosorption of Cr(VI) on the extracellular polymeric substances has
removal of cadmium via passive surface binding have been studied as been reported as the major mechanism contributing to the chromium
well. Dirbaz and Roosta have reported Parachlorella sp. as a better bioremediation by Phaeodactylum tricornutum and Navicula pelliculosa
biosorbent for Cd(II) compared to Nannochloropsis sp., Scenedesmus sp. (Hedayatkhah et al., 2018). Organelles, granules and cytosolic heat-
and Spirulina sp. due to the presence of a large amount of acidic func- stable peptides and proteins accumulated the majority of Cr(VI) inside
tional groups such as carboxylic acid. Functional groups such as eOH, microalgal cells (Aharchaou et al., 2017). Other than extracellular ad-
eNH, C]O of amide and CeO of alcoholic groups contributes to cad- sorption and intracellular accumulation of Cr(VI), enzymatic chromium
mium biosorption (Dirbaz & Roosta, 2018). The residual lipid extracted reductase is also responsible for removal of the heavy metal ions (Lee
biomass of Chlorella sp. CHA-01, Chlamydomonas sp. TAI-03 and Coe- et al., 2017; Yen et al., 2017). Furthermore, Cr(VI) also interacts with
lastrum sp. PTE-15 with broken cell wall have been investigated for the the electron donor of reducing agents (such as hydroxyl group and
biosorption of Cd2+ and suggested that the intracellular cell wall secondary alcohol groups) on the biomass surface favourably at acidic
components also contribute in heavy metal binding (Zheng et al., pH and is reduced to Cr(III), which subsequently binds to the negatively
2016). Freeze-dried biomass of Chlorella minutissima UTEX2341 was a charged functional groups (such as sulfonate group and carboxyl
superior biosorbent compared to growing algae (Yang et al., 2015). group). The reduction of Cr(VI) anionic species to Cr(III) is as following
Nostoc commune and Chlamydomonas angulosa microalgae-cellulose (Gagrai et al., 2013):
composite aerogel beads have been utilized for Cd2+ removal with at
HCrO4 + 6H+ + 3e− ⇌ Cr3+ + 4H2O E0 = 1.33 V vs. NHE
least 63.1% adsorption efficiency up to 50 ppm Cd2+ at pH 6 (Hwang
et al., 2018). Most of the literature reported that the adsorption of Cd2+ HCrO4− + 7H+ + 3e− ⇌ Cr3+ + 4H2O E0 = 1.35 V vs. NHE
followed Langmuir isotherm, suggesting that the predominant sorption
step is a single surface reaction at constant adsorption energy (Dirbaz & CrO42− + 8H+ + 3e− ⇌ Cr3+ + 4H2O E0 = 1.48 V vs. NHE
Roosta, 2018; Jena et al., 2015; Shen et al., 2017; Zhang et al., 2016; The reduction of hexavalent to trivalent chromium was even ob-
Zheng et al., 2016). Table 2 summarizes the biosorption of cadmium served in non-living microalgae biomass, which might due to the re-
(Cd) by different microalgae strains. lease of glutathione (Yen et al., 2017). Some Cr(III) complexes are re-
leased into the medium due to complex formation with adjacent
3.3. Chromium (Cr) negatively charged functional groups, ion exchange with competitor
ions (such as Ca2+, Mg2+ and Na+) as well as electrostatic repulsion
Chromium compounds have industrial applications in dyes, paint, between Cr(III) complex and positively charged functional groups
ink and pigment manufacturing, leather tanning, metal electroplating, (Pagnanelli et al., 2013; Shen et al., 2019). HPO42− and H2PO4 ions
production of steel and alloys as well as wood preservation (Gupta & significantly inhibited the adsorption of Cr(VI), while common back-
Rastogi, 2008). Chromium has several oxidation states ranging from ground anions have a minor detrimental effect on Cr(VI) reduction in
+2 to +6, with the two most common and stable states are trivalent Cr the order of SO42− < Cl− < NO3− (Gagrai et al., 2013; Singh et al.,
(III) and hexavalent Cr(VI) (Shokri Khoubestani et al., 2015). Cr(VI) is 2012).
much more toxic than Cr(III) as the former can penetrate cell membrane Many species of microalgae can tolerate and accumulate high con-
without difficulty and disturb its integrity, while the cell membrane is centrations of Cr(VI). Chlorella sorokiniana can tolerate up to 100 ppm
almost impermeable to the latter. Cr(VI) shows toxicity even at very Cr(VI) for three days and achieved removal efficiency up to 99.7% after
low concentration in parts per billion (ppb) due to its carcinogenic and 24 h contact time (Husien et al., 2019). Navicula subminuscula can tol-
mutagenic properties (Pradhan et al., 2019). Furthermore, Cr(VI) is erate high concentration of Cr(VI) up to 10 mg/L and 4 mg/L respec-
highly water-soluble and have strong oxidizing properties to damage tively, in laboratory conditions and natural chromium-containing
genetic materials and modify the DNA transcription process (Gokhale water. The microalgae can almost completely remove the heavy metal
et al., 2008; Shen et al., 2019). Ingestion of Cr(VI) can cause lung, skin up to 98% in culture containing 20 mg/L Cr(VI) (Cherifi et al., 2016).
and stomach cancer as well as chronic bronchitis, epigastric pain, liver Phaeodactylum tricornutum and Navicula pelliculosa can tolerate Cr(VI)
damage, kidney problem, tissue neurosis, ulcers or irritation in diges- up to 1 mg/L and simultaneously accumulated lipids at optimized
tive system, internal haemorrhage, emphysema and DNA impairment conditions of 55 µmol/m2s light intensity and 3 mM sodium nitrate
by interference with DNA polymerase (Daneshvar et al., 2019; concentration (Hedayatkhah et al., 2018). Pseudanabeane mucicola and
Kayalvizhi et al., 2015). Cr(III) and Cr(VI) occurs as cationic and an- Pediastrum duplex can tolerate Cr(VI) up to 1.936 and 0.224 g/L, re-
ionic species in the solution respectively, while H2CrO4 is predominant spectively, and the former has a Cr removal efficiency of 71% (Dao
at pH < 1, HCrO4− dominates between pH 1 and 6, CrO42− prevails at et al., 2018). Four microalgae presented different tolerance when
pH > 7 (Gagrai et al., 2013).

Table 2
Performance of biosorption of cadmium (Cd) by different microalgae strains.
Microalgae strains Temp (°C) Optimal pH Initial metal conc. Biomass conc. Time Max. sorption Removal efficiency Reference
(mg/L) (g/L) (min) (mg/g) (%)

Lipid-extracted Chlamydomonas 30 7.5 – 1 60 23.3 – (Zheng et al., 2016)


sp.
Lipid-extracted Chlorella sp. 30 7.5 – 1 60 25.5 –
Lipid-extracted Coelastrum sp. 30 7.5 – 1 60 32.8 –
Chlorella minutissima 28 6 – 4 20 303 – (Yang et al., 2015)
Immobilized Chlorella spa – 6 10 1.3 – 15.51 92.5 (Shen et al., 2018)
Parachlorella sp. 35 7 100 1 – 96.2 – (Dirbaz & Roosta,
2018)
Scenedesmus-24 – 6 200 1.5 – 48.4 60.5 (Jena et al., 2015)

a
A complex of water-hyacinth leaf derived biochar pellets immobilized with Chlorella sp.

5
Y.K. Leong and J.-S. Chang Bioresource Technology 303 (2020) 122886

cultivated with Cr(VI) in the order of Lyngbya sp. > Chlorella sp. > performance compared to untreated non-living consortia and live
Scenedesmus dimorphus > Oscillatoria sp. (Nath et al., 2017). Also, consortia cells (Nath et al., 2017). In addition, immobilization of mi-
Scenedesmus acutus and Chlorella vulgaris can tolerate Cr(VI) up to 15 croalgae biomass on PVA (Ardila et al., 2017) and calcium alginate
and 45 mg/L, respectively (Travieso et al., 1999). beads (Kwak et al., 2015; Singh et al., 2012) were performed for the
For Cr(VI) removal using living Anabaena, Oscillatoria, Phormidium purpose of easier harvesting. However, the harsh process of im-
and Spirogyra sp., the functional groups responsible for metal ion mobilization damaged the microalgae cell wall and resulted in reduced
binding on microalgal cell wall were carboxylate, ester and hydroxyl heavy metal removal efficiency (Ardila et al., 2017). Among the dif-
groups (Balaji et al., 2016b). The growth rate of microalgae decreased ferent forms of dried S. quadricauda (powder, pellet and biochar), mi-
with increasing chromium concentration due to delay of exponential croalgal biochar synthesized at 500 °C showed 100% removal efficiency
growth and reduction of cell division caused by heavy metal stress. up to 10 mg/L Cr(VI) (Daneshvar et al., 2019). For cost-effective Cr(VI)
Similarly, microalgae also experience a reduction in total protein con- bioremediation, lipid-extracted spent biomass (Nithya et al., 2019) and
tent with increase in heavy metal concentration, due to the accumu- β-carotene pigment-extracted biomass of Spirulina plantesis (Gokhale
lation of NADH and H+ caused by decreasing activity of the electron et al., 2008) were utilized for Cr(VI) removal. The pigment extracted
transport system. The microalgae Oscillatoria sp. is a better candidate biomass of S. plantesis demonstrated higher maximum biosorption ef-
for phycoremediation of chromium, with better biomass growth and ficiency of 86.2% compared to 73.6% of fresh biomass, which might
biosorption efficiency as well as higher antioxidant activity (Balaji due to increased surface area of biomass caused by cell membrane
et al., 2016a; Balaji et al., 2016b). Jacome-Pilco and colleagues have rupture during pigment extraction process (Gokhale et al., 2008).
studied the removal of Cr(VI) in continuous cultivation system using Ion exchange mechanism is the major contributor for chromium
Scenedesmus incrassatulus grown in a split-cylinder internal-loop airlift adsorption by microalgae where heavy metal ions carrying positive or
photobioreactor. The culture reached steady state after 14 days; uptake negative charge adsorbed on the oppositely charged biomass surface or
and removal efficiency of Cr(VI) achieved were 1.7 mg/g and 43.5%, indirect reduction followed by subsequent adsorption (Kwak et al.,
respectively (Jácome-Pilco et al., 2009). Shen and co-workers have 2015). Cr(VI) removal can also be explained by an adsorption-coupled
studied Cr(VI) removal potential of three heterotrophic microalgae reduction mechanism where Cr(VI) is removed by direct reduction of
strains namely, Botryococcus sp. NJD-1, Scenedesmus sp. NJD-6 and the biomass electron donor (Park et al., 2007). The functional groups
Chlorella sp. NJD-9 and the strain NJD-1 showed higher tolerance to Cr such as aldehydes, alkyl chains, amide, amine, alcohols/phenols, car-
(VI) compared to the other two strains. Cultivating at 3% v/v of sodium boxylic, ester, organic halide compounds, phosphate, sulfoxide and
acetate and 5 mg/L Cr(VI), the microalgae achieved removal efficiency aliphatic organic chains of cellulose were identified as functional
of 94.2%, 66.9%, 99.2% and 98.2% for Cr(VI), NO3, PO4 and TOC re- groups for the biosorption of chromium. The microalgae biosorbents
spectively. The functional groups such as –OH, –NH, aliphatic C–H, generally has maximum Cr(III) removal efficiency at pH 6, due to the
amino, carboxyl and secondary alcohol groups participated as electron increase of species such as CrOH2+ and CrOH2+ which have stronger
donors in the reduction of Cr(VI) to Cr(III) up to 87.2% (Shen et al., binding affinity to negative functional groups. On the contrary, the
2019). maximum removal of Cr(VI) was achieved at pH 1–2, due to formation
On the other hand, literature has reported the utilization of non- of anionic species of Cr(VI) such as hydrogen chromate (HCrO4−),
living biomass of microalgae as biosorbent for the removal of Cr(III) chromate (CrO42−), dichromate (Cr2O72−) and tetraoxyhydrochromate
and Cr(VI) which summarized in Table 3. Pre-treatment techniques (HCrO4−) which can be adsorbed on the protonated binding sites.
such as methylation (Finocchio et al., 2010) and treatment with NaOH Furthermore, the biomass surface consisted of high levels of hydronium
and SDS (Nath et al., 2017) were performed to improve Cr(VI) removal ions at acidic conditions which promoted interaction between Cr(VI)
by microalgal biomass. Methylated S. platensis biomass demonstrated and microalgae binding sites (Kayalvizhi et al., 2015). The adsorption
satisfactory removal efficiency of > 80% for Cr(VI) up to 25 mg/L efficiency of chromium by microalgal biosorbents increased with in-
with 2–4 g/L biomass (Finocchio et al., 2010). Pre-treated consortia of creasing temperature (Kayalvizhi et al., 2015; Nithya et al., 2019;
Scenedesmus dimorphus, Chlorella sp., Oscillatoria sp., and Lyngbya sp. Pradhan et al., 2019). Sibi also reported that Cr(VI) removal of C. vul-
with 0.1 N NaOH and 0.01% SDS both showed better Cr(VI) removal garis increased with increasing electrical conductivity up to 2.9 mS/cm,

Table 3
Performance of biosorption of chromium (Cr(VI)) by different microalgae strains.
Microalgae strain Temp (°C) Optimal pH Initial metal Biomass conc. Time (min) Max. sorption Removal ReferenceS
conc. (mg/L) (g/L) (mg/g) efficiency (%)

Spirulina platensis 60 1 500 – 90 59.6 – (Nithya et al., 2019)(Nithya


Lipid-extracted Spirulina 60 1 500 – 90 45.5 – et al., 2019)
platensis
Immobilized Spirulina 25 2 250 1 – 49 19.6 (Kwak et al., 2015)
platensisa
Methylated Spirulina platensis 25 – 18–25 2–4 120–240 16.7 > 80 (Finocchio et al., 2010)
Spirulina platensis 25 1.5 250 1 600 148.64 59.5 (Gokhale et al., 2008)
Pigment-extracted Spirulina 25 1.5 250 1 600 174.64 69.9 (Gokhale et al., 2008)
platensis (Gokhale et al., 2008)
Chlorella vulgaris 25 1.5 250 1 600 140 56
Chlorella vulgaris 25 2 147 1 240 63.2 43 (Sibi, 2016)
Scenedesmus quadricauda 25 6 for Cr(III) 100 2 120 – 98.3 for Cr(III) (Shokri Khoubestani et al.,
1 for Cr(VI) 47.6 for Cr(VI) 2015)
Scenedesmus quadricauda 22 2 10 2 240 – 100 (Daneshvar et al., 2019)
biochar
Rhizoclonium hookeri – 2 1000 1 45 67.3 6.7 (Kayalvizhi et al., 2015)
Immobilized Chlorella 30 2 100 60% (w/v) 2160 57.33 99.7 (Singh et al., 2012)
minutissimab

a
Spirulina platensis beads prepared with 1 M LiCl/DMSO.
b
Chlorella minutissima immobilized in calcium alginate beads.

6
Y.K. Leong and J.-S. Chang Bioresource Technology 303 (2020) 122886

but decreased sharply at the electrical conductivity of 3.8 mS/cm due to technology (Gan et al., 2019). Table 4 summarizes the biosorption of
competition between anionic species of Cr(VI) with chloride ion for lead (Pb) by different microalgae strains.
binding sites and inhibition of cell membrane permeability by NaCl
(Sibi, 2016).
3.5. Mercury (Hg)

3.4. Lead (Pb) Mercury has been primarily released by industrial activities such as
mining, smelting, waste incineration and coal combustion in gaseous
Lead has various applications in paint, battery making, cosmetics, and aqueous form (Peng et al., 2017). Mercury and its compounds are
weaponries and building materials. Being one of the most toxic heavy considered as one of the most toxic and hazardous heavy metals in the
metal, lead seriously endangers the safety of human health and aquatic environment as mercury can be converted into the potent neurotoxin
organisms. Other than causing serious physical and mental health methyl-mercury (MeHg). Mercury toxicity is mainly associated with its
problems in children, Pb2+ can cause anaemia, brain disease, damage capability to penetrate the blood-brain barrier, interference with the
to the central nervous system, dementia, kidney malfunction, re- uptake of essential metals, modification of cell redox status, and the
production abnormality and even death (Das et al., 2016). disruption of proteins and metal thiolate bonds (Huang et al., 2006;
Literature has reported the utilization of Spirulina platensis Rezaee et al., 2006). They can cause adverse health issues such as an-
(Malakootian et al., 2016), Chlamydomonas reinhardtii (Bayramoğlu tibiotic resistance, mental retardation, reproductive disturbance and
et al., 2006), Phormidium sp. (Das et al., 2016), Rhizoclonium hookeri others. Bioaccumulation of mercury is similar to hydrophobic organic
(Suganya et al., 2016) as well as Chaetoceros and Chlorella sp. compounds, rather than inorganic metal ions, as “biomagnification” of
(Molazedah et al., 2015) as the biosorbent for the removal of lead. Das mercury happens at every level of aquatic food chain (Mason et al.,
and co-workers have further studied semi-packed bed adsorption of 1996).
Pb2+ and found that bed height and flow rate were the significant Microalgae have the ability to bio-transform acid reducible Hg2+
parameters in the semi-batch biosorption process (Das et al., 2016). The into elemental Hg0 and metacinnabar (β-HgS) to varying degrees (Kelly
functional groups such as acyl-amino, amide, amine, carbonyl, car- et al., 2007). After enzymatic reduction of Hg2+ to Hg0 catalysed by
boxyl, hydroxyl, phenols and phosphate were identified for the bio- mercuric reductase, the volatile Hg0 is removed by both biological and
sorption of Pb2+. The adsorption of lead ions decreased significantly at non-biological volatilization and most of the remaining unreduced
low pH due to electrostatic repulsion caused by high positive charge Hg2+ is converted into β-HgS. Majority of Hg0 volatilization happens
density on the binding sites and competition with H+ for occupancy of rapidly, within 20 min to few hours, as shown in microalgae such as
sorption sites. The biosorbent generally has the maximum efficiency of Selenastrum minutum, Chlorella fusca var. fusca, diatom Navicula pellicosa
Pb2+ removal at pH 5–6, due to the increase of species such as Pb2+ and thermophilic alga Galdiera sulphuraria. Non-biological volatiliza-
and Pb(OH)+. The formation of Pb(OH)2 which precipitated was ob- tion of mercury only happened under light illumination, while biolo-
served at pH > 6 (Akhtar et al., 2004). Ion exchange mechanism can gical volatilization was light-independent, but depended on metal
be employed to explain the adsorption of cationic species of Pb2+ on concentration and cell density (Devars et al., 2000). Accelerated vola-
Chlorella sp. surface. The organometallic complex formation model is as tilization was shown in microalgae pretreated with 5 mM dimethylfu-
follows (Molazedah et al., 2015): marate (Kelly et al., 2007). Genetic engineering has been performed to
express mercuric reductase (MerA) from Bacillus megaterium B1 in
2(-R-OH) + Pb2+ → 2(RO)Pb + 2H+ Chlorella sp. DT. The transgenic strains have improved Hg(II) removal
ability, up to two fold compared to control and experienced a reduced
-R-OH + Pb(OH)+ → (-RO)PbOH + H+
level of oxidative stress (Huang et al., 2006). In addition, other po-
The adsorption enthalpy value and enthalpy change showed that the tential routes of Hg(II) removal are thiol chelation and bio-methylation
adsorption of Pb2+ by microalgae biomass is spontaneous, endothermic to methylmercury (MeHg). However, it was reported that unlike other
and chemical in nature. Most studies agreed that the biosorption of heavy metals, accumulation of mercury poorly induces phytochelatin
Pb2+ by microalgae followed a pseudo-second order kinetic model (Das synthesis (Devars et al., 2000). Growth photoperiods have reported to
et al., 2016; Malakootian et al., 2016; Molazedah et al., 2015; Suganya influence the composition of binding ligands and its complexation with
et al., 2016). This suggested that the chemical sorption which involves Hg. Hg-binding ligands are more homologous and aromatic in nature at
ion exchange and electron sharing between Pb2+ and the biomass longer light exposure period, while formation of smaller, more aliphatic
might be the rate-limiting step. Hg-ligand complexes rich in thiols and sulphur was observed at darker
Upon the exposure to lead, Scenedesmus incrassatulus showed a de- growth conditions (Mangal et al., 2019).
creased peripheral cell type “incrassatulus” and increased morphology A novel pretreatment technique, polyelectrolyte self-assembly to-
type “obliquus”. Therefore, peripheral morphology of S. incrassatulus gether with biomimetic mineralization was used to pretreat Chlorella
can be employed as an effective bio-indicator for the detection of lead vulgaris for Hg2+ removal. The mineralized microalgae cells were
pollution (Batsalova et al., 2017). Similarly, marine microalgae Nitzchia wrapped in a coating of polyelectrolytes and a transparent amorphous
closterium has become a potential candidate for rapid, simple and ef- calcium phosphate mineral layer doped with sulphur atoms. The mi-
fective detection of lead toxicity utilizing chlorophyll fluorescence neral layer serves as a protection layer for microalgae from heavy

Table 4
Performance of biosorption of lead (Pb) by different microalgae strains.
Microalgae strains Temp (°C) Optimal pH Initial metal Biomass conc. Time Max. sorption Removal Reference
conc. (mg/L) (g/L) (min) (mg/g) efficiency (%)

Chaetoceros sp. 25 6 20 1.5 180 8 60 (Molazedah et al.,


Chlorella sp. 25 6 20 1.5 180 10.4 78 2015)
Immobilized Chlamydomonas 25 6 500 – 120 308.7 – (Bayramoğlu et al.,
reinhardtiia 2006)
Phormidium sp. 25 5 10 4 40 2.305 92.2 (Das et al., 2016)
Rhizoclonium hookeri 40 4.5 – – – 81.7 – (Suganya et al., 2016)

a
Chlamydomonas reinhardtii immobilized in calcium alginate beads.

7
Y.K. Leong and J.-S. Chang Bioresource Technology 303 (2020) 122886

metals poisoning and the microalgae can grow well in high con- 5. Future prospects
centration of mercury up to 100 µg/L with no obvious reduction in lipid
yield. Raw C. vulgaris has the optimal pH of 5.5 for the removal of Hg2+ Considering the multiple advantages provided by microalgae for
up to 62.85%, as Hg2+ compete with H+ for binding sites at lower pH, heavy metal bioremediation, the future prospects of large scale appli-
while complex formation of (Hg)OH2 at higher pH significantly de- cation look promising by taking into accounts few considerations.
terred the binding of Hg2+ with organic functional groups. On the other Firstly, screening and choosing the suitable microalgae strains is an
hand, the biomimetic mineralized C. vulgaris has enhanced adsorption important step in bioremediation of heavy metals wastewater. Other
efficiency with increasing pH up to 94.7% at pH 7, due to higher degree than having high tolerance to heavy metal pollutants with fast and
of crystallinity and improved loading of S atoms on the mineralized stable growth, it is beneficial if the microalgae has the following at-
layer (Peng et al., 2017). tributes: (i) able to accumulate high content of lipids and other valuable
Integration of membrane filtration and activated sludge utilizing co-products, (ii) high CO2 sequestration ability and low nutrient re-
membrane bioreactor systems for wastewater treatment has gained a lot quirements, (iii) resistant to predation by grazers and robustness to-
of attention due to its high efficiency. Microalgae dynamic membrane ward existence of other microorganism species, and (iv) ability to self-
formed using C. vulgaris powder was utilized in dynamic membrane flocculate for cost-effective cell harvesting. Therefore, genetic, meta-
bioreactor for the bioremediation of Hg2+ from synthetic dental was- bolic and molecular engineering to increase adaptive ability, specificity
tewater with the advantages of promoted removal yield as well as re- and robustness of microalgae strains is a potential research area.
duction in fouling and expensive cost of membrane recovery. The dy- Similarly, continuous research for better understanding of the un-
namic membrane bioreactor demonstrated a higher Hg2+ removal derlying mechanisms as well as development of more comprehensive
efficiency compared to control membrane bioreactor at 300–800 ppb equilibrium and kinetic model for heavy metals biosorption and
concentrations (Fard & Mehrnia, 2017). bioaccumulation by microalgae is also crucial. Recently, techniques
The development of a rapid and sensitive biological monitoring such as whole-cell immobilization, pelletisation and microalgal biofilm
technology for heavy metal pollutants is of utmost importance for the for heavy metals removal and recovery have gained much attention
protection of aquatic environment. Other than significant increase in because of their potential in industrial applications. Also, strategies
SOD gene expression levels and activity as well as chlorophyll a con- such as surface and chemical modifications techniques as well as in-
tent, the frustule of Halamphora veneta deformed and increased in size tegration with other heavy metal removal techniques for microalgae
when exposed to mercury. Therefore, SOD gene expression levels and biomass can help to improve heavy metals removal efficiency.
activity, changes in frustule morphology and chlorophyll a content of Microalgae biomass harvesting has always been a challenging task as
Halamphora veneta are promising candidates for mercury toxicological microalgae are minute in size and their negatively charged surface keep
assessment in aquatic habitats (Mu et al., 2017). Table 5 summarizes them in stable dispersed state. Innovation of conventional harvesting
the biosorption of mercury (Hg) by different microalgae strains. technologies and development of new techniques are essential to reduce
the overall operating cost.
On the other hands, heavy metals have both enhancement and in-
4. Challenges hibition effect on microalgae growth as well as accumulation of lipid
and other valuable co-products. Although usually having inhibition
Although microalgae offer a number of attractive benefits for heavy effect on microalgae growth, some heavy metals such as Al, As, Cd Co,
metal bioremediation, it still faces some challenges for widespread Pb, elements from the lanthanide group, metallic nanoparticles and
applications, such as contamination by other microorganisms, nutrient others at certain concentration demonstrated a positive influence on the
variability, high content of TSS and high turbidity, microalgae biomass growth of specific microalgae strains as reported in a recent review
harvesting, difficulty in downstream processing and others. Table 6 (Miazek et al., 2015). Similarly, certain heavy metals also played a
summarizes the issues that could challenge the feasibility of using mi- positive role in the accumulation of lipids (such as As, Cd, Cu, Ni),
croalgae-based biosorbents for heavy metal removal as well as the exopolymers (such as Ag, Cd, Co, Cu), pigments (such as As, Cd, Cu, Fe,
proposed strategies that could solve the problems raise when applying Ni and Te), phytochelatin (such as As, Cd, Cu, Pb) and phytohormones
this technology. From the information provided in Table 6, heavy metal (such as Cd, Cu, Pb) due to the inducing effect by metal stress. In
removal by microalgae seems to be a promising and efficient alternative particular, heavy metal stress can be exploited to modify microalgal
to conventional methods when an appropriate pretreatment step is fatty acids composition in order to produce biodiesel with desirable
employed or when it is integrated with other technologies. Still, more properties and quality (Miazek et al., 2015). However, the presence of
powerful microalgal strains should be isolated and better treatment heavy metals might interfere with downstream processing of these va-
processes should be developed to further improve the efficiency of luable products. Therefore, more research work should be done on
heavy metal remediation and to reduce the operation costs. The future upstream cultivation and downstream purification process to achieve
prospects are thus discussed in detail below. satisfactory simultaneous heavy metal bioremediation and value-added
product formation.

Table 5
Performance of biosorption of mercury (Hg) by different microalgae strains.
Microalgae strains Temp (°C) Optimal pH Initial metal Biomass conc. Time Max. sorption Removal Reference
conc. (mg/L) (g/L) (min) (mg/g) efficiency (%)

Immobilized Chlamydomonas 25 6 500 – 120 106.6 – (Bayramoğlu et al.,


reinhardtiia 2006)
Chlorella sp. DT 30 – 8 0.3 120 3.33 12.5 (Huang et al., 2006)
Transgenic Chlorella sp.b 30 – 8 0.3 120 7.33 27.5
Chlorella vulgaris 20 5 48 2 120 17.49 72.9 (Solisio et al., 2019)
Scenedesmus obtusus XJ-15 25 5 20 0.125 180 – – (Huang et al., 2019)
Spirogyra sp. 4 4 1 3 30 0.253 76 (Rezaee et al., 2006)

a
Chlamydomonas reinhardtii immobilized in calcium alginate beads.
b
Transformed with the Bacillus megaterium strain MB1 merA gene.

8
Y.K. Leong and J.-S. Chang Bioresource Technology 303 (2020) 122886

Table 6
Challenges and proposed strategies of heavy metal removal by microalgae.
Challenges Proposed strategies

Raw heavy metal wastewater consists of bacteria, fungus and other microorganisms, which
compete for nutrients and possibly dominating because of their relatively higher
• Pretreatment techniques (such as acidification, autoclaving, chlorination by
bleach, exposure to high level of ammonia, filtration, ozonation and UV-
growth rate irradiation)
• Utilization of microalgae-bacteria, microalgae-fungus and others consortium
Nutrient variability (such as excessive or deprivation in carbon, nitrogen and phosphorus • Nutrient supplementation using waste nutrient rich materials
content) • Dilute high strength wastewater with low nutrient wastewater
• Select the appropriate microalgae strain that can adapt to the specific wastewater
• Screening
wastewater
and isolation of indigenous microalgae species from heavy metal

Wastewater might contain huge amount of TSS and has high turbidity that reduce light
penetration and hinder microalgae growth
• gravity)
Pretreatment techniques (such as adding flocculants and sedimentation by

• Introduce turbulence such as in raceway ponds to increase light exposure


Microalgae biomass harvesting • Combination of different harvesting techniques and integration with other
technologies (such as membrane technology and electrokinetics technologies)
• Exploring the potential of novel natural organic flocculants, coagulants, magnetic
and nanoparticles separator
• Strategies such as immobilization, microalgal biofilm, pelletisation, co-cultivation
with flocculating microorganisms
Presence of heavy metal remaining in microalgal cell might interfere with downstream • Pretreatment techniques (such as recovery and regeneration of heavy metals)
processing of valuable product • Integration and hybrid downstream purification strategies to remove heavy metals
6. Conclusions Mater. 289, 38–45.
Ardila, L., Godoy, R., Montenegro, L., 2017. Sorption capacity measurement of chlorella
vulgaris and scenedesmus acutus to remove chromium from tannery waste water. IOP
Microalgae strains demonstrated varied tolerance and response as Conf. Series: Earth Environ. Sci. 83, 012031.
well as bioaccumulation capability towards heavy metals. Different Arora, N., Dubey, D., Sharma, M., Patel, A., Guleria, A., Pruthi, P.A., Kumar, D., Pruthi,
functional groups as well as proteins and peptides are responsible for V., Poluri, K.M., 2018. NMR-based metabolomic approach to elucidate the differ-
ential cellular responses during mitigation of arsenic(III, V) in a green microalga. ACS
metal-binding. Mechanisms such as extracellular adsorption, reduction, Omega 3 (9), 11847–11856.
volatilization, complex formation, ion exchange, intracellular accumu- Arora, N., Gulati, K., Patel, A., Pruthi, P.A., Poluri, K.M., Pruthi, V., 2017. A hybrid ap-
lation, chelation and bio-methylation are involved in bioremediation proach integrating arsenic detoxification with biodiesel production using oleaginous
microalgae. Algal Res. 24, 29–39.
and biosorption of heavy metals. More efforts in the areas of genetic
Bahar, M.M., Megharaj, M., Naidu, R., 2016. Influence of phosphate on toxicity and
engineering, immobilization techniques, pretreatment strategies and bioaccumulation of arsenic in a soil isolate of microalga Chlorella sp. Environ. Sci.
integration with other technologies are required to fully explore the Pollut. Res. 23 (3), 2663–2668.
Baker, J., Wallschläger, D., 2016. The role of phosphorus in the metabolism of arsenate by
potential of microalgae in heavy metal phycoremediation and si-
a freshwater green alga, Chlorella vulgaris. J. Environ. Sci. 49, 169–178.
multaneously value-added products formation. Balaji, S., Kalaivani, T., Shalini, M., Gopalakrishnan, M., Rashith Muhammad, M.A.,
Rajasekaran, C., 2016a. Sorption sites of microalgae possess metal binding ability
CRediT authorship contribution statement towards Cr(VI) from tannery effluents—a kinetic and characterization study. Desalin.
Water Treat. 57 (31), 14518–14529.
Balaji, S., Kalaivani, T., Sushma, B., Pillai, C.V., Shalini, M., Rajasekaran, C., 2016b.
Yoong Kit Leong: Investigation, Writing - original draft. Jo-Shu Characterization of sorption sites and differential stress response of microalgae iso-
Chang: Supervision, Conceptualization, Writing - review & editing. lates against tannery effluents from ranipet industrial area—an application towards
phycoremediation. Int. J. Phytorem. 18 (8), 747–753.
Batsalova, T., Teneva, I., Belkinova, D., Stoyanov, P., Rusinova-Videva, S., Dzhambazov,
Declaration of competing Interest B., 2017. Assessment of cadmium, nickel and lead toxicity by using green algae
scenedesmus incrassatulus and human cell lines: potential in vitro test-systems for
monitoring of heavy metal pollution. Toxicol. Forens. Med. 2 (2), 63–73.
The authors declare that they have no known competing financial Bayramoğlu, G., Tuzun, I., Celik, G., Yilmaz, M., Arica, M.Y., 2006. Biosorption of mer-
interests or personal relationships that could have appeared to influ- cury(II), cadmium(II) and lead(II) ions from aqueous system by microalgae
ence the work reported in this paper. Chlamydomonas reinhardtii immobilized in alginate beads. Int. J. Miner. Process. 81
(1), 35–43.
Birungi, Z.S., Chirwa, E.M.N., 2015. The adsorption potential and recovery of thallium
Acknowledgements using green micro-algae from eutrophic water sources. J. Hazard. Mater. 299, 67–77.
Bodin, H., Asp, H., Hultberg, M., 2017. Effects of biopellets composed of microalgae and
fungi on cadmium present at environmentally relevant levels in water. Int. J.
The authors gratefully acknowledge financial support received from
Phytorem. 19 (5), 500–504.
Taiwan’s Ministry of Science and Technology under grant number Cameron, H., Mata, M.T., Riquelme, C., 2018. The effect of heavy metals on the viability
MOST 109-3116-F-006 -016 -CC1; 108-2621-M-006-020, 108-2218-E- of Tetraselmis marina AC16-MESO and an evaluation of the potential use of this
029-002-MY3, and 107-2221-E-006-112-MY3. microalga in bioremediation. PeerJ 6, e5295.
Cherifi, O., Sbihi, K., Bertrand, M., Cherifi, K., 2016. The siliceous microalga Navicula
subminuscula (Manguin) as a biomaterial for removing metals from tannery effluents:
References a laboratory study. J. Mater. Environ. Sci. 8 (3), 884–893.
Chia, M.A., Lombardi, A.T., Melão, M.d.G.G., Parrish, C.C., 2013. Lipid composition of
Chlorella vulgaris (Trebouxiophyceae) as a function of different cadmium and
Abinandan, S., Subashchandrabose, S.R., Venkateswarlu, K., Perera, I.A., Megharaj, M., phosphate concentrations. Aquat. Toxicol. 128–129, 171–182.
2019. Acid-tolerant microalgae can withstand higher concentrations of invasive Daneshvar, E., Zarrinmehr, M.J., Kousha, M., Hashtjin, A.M., Saratale, G.D., Maiti, A.,
cadmium and produce sustainable biomass and biodiesel at pH 3.5. Bioresour. Vithanage, M., Bhatnagar, A., 2019. Hexavalent chromium removal from water by
Technol. 281, 469–473. microalgal-based materials: adsorption, desorption and recovery studies. Bioresour.
Aharchaou, I., Rosabal, M., Liu, F., Battaglia, E., Vignati, D.A.L., Fortin, C., 2017. Technol. 293, 122064.
Bioaccumulation and subcellular partitioning of Cr(III) and Cr(VI) in the freshwater Dao, T.-S., Le, N.-H.-S., Vo, M.-T., Vo, T.-M.-C., Phan, T.-H., Bui, T.-N.-P., 2018. Growth
green alga Chlamydomonas reinhardtii. Aquat. Toxicol. 182, 49–57. and metal uptake capacity of microalgae under exposure to chromium. J. Vietnam.
Akhtar, N., Iqbal, J., Iqbal, M., 2004. Enhancement of Lead(II) Biosorption by Microalgal Environ. 9 (1), 38–43.
Biomass Immobilized onto Loofa (Luffa cylindrica) Sponge. Eng. Life Sci. 4 (2), Das, D., Chakraborty, S., Bhattacharjee, C., Chowdhury, R., 2016. Biosorption of lead ions
171–178. (Pb2+) from simulated wastewater using residual biomass of microalgae. Desalin.
Alam, M.A., Wan, C., Zhao, X.-Q., Chen, L.-J., Chang, J.-S., Bai, F.-W., 2015. Enhanced Water Treat. 57 (10), 4576–4586.
removal of Zn2+ or Cd2+ by the flocculating Chlorella vulgaris JSC-7. J. Hazard. Devars, S., Avilés, C., Cervantes, C., Moreno-Sánchez, R., 2000. Mercury uptake and

9
Y.K. Leong and J.-S. Chang Bioresource Technology 303 (2020) 122886

removal by Euglena gracilis. Arch. Microbiol. 174 (3), 175–180. Li, B., Zhang, T., Yang, Z., 2019. Immobilizing unicellular microalga on pellet-forming
Dirbaz, M., Roosta, A., 2018. Adsorption, kinetic and thermodynamic studies for the filamentous fungus: Can this provide new insights into the remediation of arsenic
biosorption of cadmium onto microalgae Parachlorella sp. J. Environ. Chem. Eng. 6 from contaminated water? Bioresour. Technol. 284, 231–239.
(2), 2302–2309. Malakootian, M., Khodashenas Limoni, Z., Malakootian, M., 2016. The efficiency of lead
Doshi, H., Ray, A., 2009. Live and dead spirulina SP. to remove arsenic (V) from wate. Int. biosorption from industrial wastewater by micro-alga spirulina platensis. Int. J.
J. Phytorem. 11 (1), 53–64. Environ. Res. 10 (3), 357–366.
Duruibe, J.O., Ogwuegbu, M.O.C., Egwuruhwu, J.N., 2007. Heavy metal pollution and Mangal, V., Phung, T., Nguyen, T.Q., Gueguen, C., 2019. Molecular characterization of
human biotoxic effect. Int. J. Phys. Sci. 2 (5), 112–118. mercury binding ligands released by freshwater algae grown at three photoperiods.
Fard, G.H., Mehrnia, M.R., 2017. Investigation of mercury removal by micro-algae dy- Front. Environ. Sci. 6 (155).
namic membrane bioreactor from simulated dental waste water. J. Environ. Chem. Mason, R.P., Reinfelder, J.R., Morel, F.M.M., 1996. Uptake, toxicity, and trophic transfer
Eng. 5 (1), 366–372. of mercury in a coastal diatom. Environ. Sci. Technol. 30 (6), 1835–1845.
Finocchio, E., Lodi, A., Solisio, C., Converti, A., 2010. Chromium (VI) removal by me- Mera, R., Torres, E., Abalde, J., 2016. Influence of sulphate on the reduction of cadmium
thylated biomass of Spirulina platensis: the effect of methylation process. Chem. Eng. toxicity in the microalga Chlamydomonas moewusii. Ecotoxicol. Environ. Saf. 128,
J. 156 (2), 264–269. 236–245.
Fu, F., Wang, Q., 2011. Removal of heavy metal ions from wastewaters: a review. J. Miazek, K., Iwanek, W., Remacle, C., Richel, A., Goffin, D., 2015. Effect of metals, me-
Environ. Manage. 92 (3), 407–418. talloids and metallic nanoparticles on microalgae growth and industrial product
Gagrai, M.K., Das, C., Golder, A.K., 2013. Reduction of Cr(VI) into Cr(III) by Spirulina biosynthesis: a review. Int. J. Mol. Sci. 16 (10).
dead biomass in aqueous solution: kinetic studies. Chemosphere 93 (7), 1366–1371. Molazedah, P., Khanjani, N., Rahimi, M.R., Nasiri, A., 2015. Adsorption of lead by mi-
Gan, T., Zhao, N., Yin, G., Chen, M., Wang, X., Liu, J., Liu, W., 2019. Optimal chlorophyll croalgae Chaetoceros sp. and Chlorella sp. from aqueous solution. J. Commun. Health
fluorescence parameter selection for rapid and sensitive detection of lead toxicity to Res. 4 (2), 114–127.
marine microalgae Nitzschia closterium based on chlorophyll fluorescence tech- Morelli, E., Mascherpa, M.C., Scarano, G., 2005. Biosynthesis of phytochelatins and ar-
nology. J. Photochem. Photobiol., B 197, 111551. senic accumulation in the marine microalga phaeodactylum tricornutum in response
Gokhale, S.V., Jyoti, K.K., Lele, S.S., 2008. Kinetic and equilibrium modeling of chromium to arsenate exposure. Biometals 18 (6), 587–593.
(VI) biosorption on fresh and spent Spirulina platensis/Chlorella vulgaris biomass. Mu, W., Jia, K., Liu, Y., Pan, X., Fan, Y., 2017. Response of the freshwater diatom
Bioresour. Technol. 99 (9), 3600–3608. Halamphora veneta (Kützing) Levkov to copper and mercury and its potential for
Gómez-Jacinto, V., García-Barrera, T., Gómez-Ariza, J.L., Garbayo-Nores, I., Vílchez- bioassessment of heavy metal toxicity in aquatic habitats. Environ. Sci. Pollut. Res. 24
Lobato, C., 2015. Elucidation of the defence mechanism in microalgae Chlorella (34), 26375–26386.
sorokiniana under mercury exposure. Identification of Hg–phytochelatins. Chem. Nath, A., Tiwari, P.K., Rai, A.K., Sundaram, S., 2017. Microalgal consortia differentially
Biol. Interact. 238, 82–90. modulate progressive adsorption of hexavalent chromium. Physiol. Mol. Biol. Plants
Gupta, V.K., Rastogi, A., 2008. Biosorption of lead from aqueous solutions by green algae 23 (2), 269–280.
Spirogyra species: kinetics and equilibrium studies. J. Hazard. Mater. 152 (1), Nithya, K., Sathish, A., Pradeep, K., Kiran Baalaji, S., 2019. Algal biomass waste residues
407–414. of Spirulina platensis for chromium adsorption and modeling studies. J. Environ.
Hedayatkhah, A., Cretoiu, M.S., Emtiazi, G., Stal, L.J., Bolhuis, H., 2018. Bioremediation Chem. Eng. 7 (5), 103273.
of chromium contaminated water by diatoms with concomitant lipid accumulation Pagnanelli, F., Jbari, N., Trabucco, F., Martínez, M.E., Sánchez, S., Toro, L., 2013.
for biofuel production. J. Environ. Manage. 227, 313–320. Biosorption-mediated reduction of Cr(VI) using heterotrophically-grown Chlorella
Huang, C.C., Chen, M.W., Hsieh, J.L., Lin, W.H., Chen, P.C., Chien, L.F., 2006. Expression vulgaris: active sites and ionic strength effect. Chem. Eng. J. 231, 94–102.
of mercuric reductase from Bacillus megaterium MB1 in eukaryotic microalga Papry, R.I., Ishii, K., Mamun, M.A.A., Miah, S., Naito, K., Mashio, A.S., Maki, T.,
Chlorella sp. DT: an approach for mercury phytoremediation. Appl. Microbiol. Hasegawa, H., 2019. Arsenic biotransformation potential of six marine diatom spe-
Biotechnol. 72 (1), 197–205. cies: effect of temperature and salinity. Sci. Rep. 9 (1), 10226.
Huang, R., Huo, G., Song, S., Li, Y., Xia, L., Gaillard, J.-F., 2019. Immobilization of Park, D., Lim, S.-R., Yun, Y.-S., Park, J.M., 2007. Reliable evidences that the removal
mercury using high-phosphate culture-modified microalgae. Environ. Pollut. 254, mechanism of hexavalent chromium by natural biomaterials is adsorption-coupled
112966. reduction. Chemosphere 70 (2), 298–305.
Huang, W.-J., Wu, C.-C., Chang, W.-C., 2014. Bioaccumulation and toxicity of arsenic in Peng, Y., Deng, A., Gong, X., Li, X., Zhang, Y., 2017. Coupling process study of lipid
cyanobacteria cultures separated from a eutrophic reservoir. Environ. Monit. Assess. production and mercury bioremediation by biomimetic mineralized microalgae.
186 (2), 805–814. Bioresour. Technol. 243, 628–633.
Husien, S., Labena, A., El-Belely, E.F., Mahmoud, H.M., Hamouda, A.S., 2019. Absorption Pradhan, D., Sukla, L.B., Mishra, B.B., Devi, N., 2019. Biosorption for removal of hex-
of hexavalent chromium by green micro algae Chlorella sorokiniana: live planktonic avalent chromium using microalgae Scenedesmus sp. J. Cleaner Prod. 209, 617–629.
cells. Water Pract. Technol. 14 (3), 515–529. Priatni, S., Ratnaningrum, D., Warya, S., Audina, E., 2017. Phycobiliproteins production
Hwang, K., Kwon, G.-J., Yang, J., Kim, M., Hwang, W.J., Youe, W., Kim, D.-Y., 2018. and heavy metals reduction ability of Porphyridium sp. IOP Conf. Series: Earth
Chlamydomonas angulosa (Green Alga) and Nostoc commune (Blue-Green Alga) Environ. Sci, 60.
microalgae-cellulose composite aerogel beads: manufacture, physicochemical char- Rahman, Z., Singh, V.P., 2019. The relative impact of toxic heavy metals (THMs) (arsenic
acterization, and Cd (II) adsorption. Materials (Basel, Switzerland) 11 (4), 562. (As), cadmium (Cd), chromium (Cr)(VI), mercury (Hg), and lead (Pb)) on the total
Ibuot, A., Dean, A.P., McIntosh, O.A., Pittman, J.K., 2017. Metal bioremediation by environment: an overview. Environ. Monit. Assess. 191 (7), 419.
CrMTP4 over-expressing Chlamydomonas reinhardtii in comparison to natural was- Rezaee, A., Ramavandi, B., Ganati, F., Ansari, M., Solimanian, A., 2006. Biosorption of
tewater-tolerant microalgae strains. Algal Res. 24, 89–96. mercury by biomass of filamentous algae Spirogyra species. J. Biol. Sci. 6 (4),
Jaafari, J., Yaghmaeian, K., 2019. Optimization of heavy metal biosorption onto fresh- 695–700.
water algae (Chlorella coloniales) using response surface methodology (RSM). Saavedra, R., Muñoz, R., Taboada, M.E., Vega, M., Bolado, S., 2018. Comparative uptake
Chemosphere 217, 447–455. study of arsenic, boron, copper, manganese and zinc from water by different green
Jácome-Pilco, C.R., Cristiani-Urbina, E., Flores-Cotera, L.B., Velasco-García, R., Ponce- microalgae. Bioresour. Technol. 263, 49–57.
Noyola, T., Cañizares-Villanueva, R.O., 2009. Continuous Cr(VI) removal by Sarı, A., Uluozlü, Ö.D., Tüzen, M., 2011. Equilibrium, thermodynamic and kinetic in-
Scenedesmus incrassatulus in an airlift photobioreactor. Bioresour. Technol. 100 (8), vestigations on biosorption of arsenic from aqueous solution by algae (Maugeotia
2388–2391. genuflexa) biomass. Chem. Eng. J. 167 (1), 155–161.
Jena, J., Pradhan, N., Aishvarya, V., Nayak, R.R., Dash, B.P., Sukla, L.B., Panda, P.K., Shen, L., Saky, S.A., Yang, Z., Ho, S.-H., Chen, C., Qin, L., Zhang, G., Wang, Y., Lu, Y.,
Mishra, B.K., 2015. Biological sequestration and retention of cadmium as CdS na- 2019. The critical utilization of active heterotrophic microalgae for bioremoval of Cr
noparticles by the microalga Scenedesmus-24. J. Appl. Phycol. 27 (6), 2251–2260. (VI) in organics co-contaminated wastewater. Chemosphere 228, 536–544.
Jiang, Y., Purchase, D., Jones, H., Garelick, H., 2011. Effects of arsenate (AS5+) on Shen, Y., Li, H., Zhu, W., Ho, S.-H., Yuan, W., Chen, J., Xie, Y., 2017. Microalgal-biochar
growth and production of glutathione (GSH) and phytochelatins (PCS) in Chlorella immobilized complex: a novel efficient biosorbent for cadmium removal from aqu-
vulgaris. Int. J. Phytoremediat. 13 (8), 834–844. eous solution. Bioresour. Technol. 244, 1031–1038.
Karadjova, I.B., Slaveykova, V.I., Tsalev, D.L., 2008. The biouptake and toxicity of arsenic Shen, Y., Zhu, W., Li, H., Ho, S.-H., Chen, J., Xie, Y., Shi, X., 2018. Enhancing cadmium
species on the green microalga Chlorella salina in seawater. Aquat. Toxicol. 87 (4), bioremediation by a complex of water-hyacinth derived pellets immobilized with
264–271. Chlorella sp. Bioresour. Technol. 257, 157–163.
Kayalvizhi, K., Vijayaraghavan, K., Velan, M., 2015. Biosorption of Cr(VI) using a novel Shokri Khoubestani, R., Mirghaffari, N., Farhadian, O., 2015. Removal of three and
microalga Rhizoclonium hookeri: equilibrium, kinetics and thermodynamic studies. hexavalent chromium from aqueous solutions using a microalgae biomass-derived
Desalin. Water Treat. 56 (1), 194–203. biosorbent. Environ. Prog. Sustainable Energy 34 (4), 949–956.
Kelly, D.J.A., Budd, K., Lefebvre, D.D., 2007. Biotransformation of mercury in pH-stat Sibi, G., 2016. Biosorption of chromium from electroplating and galvanizing industrial
cultures of eukaryotic freshwater algae. Arch. Microbiol. 187 (1), 45–53. effluents under extreme conditions using Chlorella vulgaris. Green Energy Environ. 1
Kwak, H.W., Kim, M.K., Lee, J.Y., Yun, H., Kim, M.H., Park, Y.H., Lee, K.H., 2015. (2), 172–177.
Preparation of bead-type biosorbent from water-soluble Spirulina platensis extracts Singh, N.K., Raghubanshi, A.S., Upadhyay, A.K., Rai, U.N., 2016. Arsenic and other heavy
for chromium (VI) removal. Algal Res. 7, 92–99. metal accumulation in plants and algae growing naturally in contaminated area of
Lee, L., Hsu, C.-Y., Yen, H.-W., 2017. The effects of hydraulic retention time (HRT) on West Bengal, India. Ecotoxicol. Environ. Saf. 130, 224–233.
chromium(VI) reduction using autotrophic cultivation of Chlorella vulgaris. Singh, S.K., Bansal, A., Jha, M.K., Dey, A., 2012. An integrated approach to remove Cr(VI)
Bioprocess Biosyst. Eng. 40 (12), 1725–1731. using immobilized Chlorella minutissima grown in nutrient rich sewage wastewater.
Levy, J.L., Stauber, J.L., Adams, M.S., Maher, W.A., Kirby, J.K., Jolley, D.F., 2005. Bioresour. Technol. 104, 257–265.
Toxicity, biotransformation, and mode of action of arsenic in two freshwater mi- Solisio, C., Al Arni, S., Converti, A., 2019. Adsorption of inorganic mercury from aqueous
croalgae (Chlorella sp. and Monoraphidium arcuatum). Environ. Toxicol. Chem. 24 solutions onto dry biomass of Chlorella vulgaris: kinetic and isotherm study. Environ.
(10), 2630–2639. Technol. 40 (5), 664–672.

10
Y.K. Leong and J.-S. Chang Bioresource Technology 303 (2020) 122886

Suganya, S., Saravanan, A., Senthil Kumar, P., Yashwanthraj, M., Sundar Rajan, P., phosphate regimes. Water Res. 47 (7), 2497–2506.
Kayalvizhi, K., 2016. Sequestration of Pb(II) and Ni(II) ions from aqueous solution Wang, Y., Zhang, C., Zheng, Y., Ge, Y., 2017. Bioaccumulation kinetics of arsenite and
using microalga Rhizoclonium hookeri: adsorption thermodynamics, kinetics, and arsenate in Dunaliella salina under different phosphate regimes. Environ. Sci. Pollut.
equilibrium studies. J. Water Reuse Desalin. 7 (2), 214–227. Res. 24 (26), 21213–21221.
Sulaymon, A.H., Mohammed, A.A., Al-Musawi, T.J., 2013a. Column biosorption of lead, Wang, Y., Zhang, C.H., Lin, M.M., Ge, Y., 2016. A symbiotic bacterium differentially
cadmium, copper, and arsenic ions onto algae. J. Bioprocess. Biotechniq. 3 (1). influences arsenate absorption and transformation in Dunaliella salina under dif-
Sulaymon, A.H., Mohammed, A.A., Al-Musawi, T.J., 2013b. Removal of lead, cadmium, ferent phosphate regimes. J. Hazard. Mater. 318, 443–451.
copper, and arsenic ions using biosorption: equilibrium and kinetic studies. Desalin. Wongrod, S., Simon, S., van Hullebusch, E.D., Lens, P.N.L., Guibaud, G., 2019. Assessing
Water Treat. 51 (22–24), 4424–4434. arsenic redox state evolution in solution and solid phase during As(III) sorption onto
Sun, J., Cheng, J., Yang, Z., Li, K., Zhou, J., Cen, K., 2015. Microstructures and functional chemically-treated sewage sludge digestate biochars. Bioresour. Technol. 275,
groups of Nannochloropsis sp. cells with arsenic adsorption and lipid accumulation. 232–238.
Bioresour. Technol. 194, 305–311. Yang, J., Cao, J., Xing, G., Yuan, H., 2015. Lipid production combined with biosorption
Suvarapu, L.N., Baek, S.-O., 2017. Determination of heavy metals in the ambient atmo- and bioaccumulation of cadmium, copper, manganese and zinc by oleaginous mi-
sphere: a review. Toxicol. Ind. Health 33 (1), 79–96. croalgae Chlorella minutissima UTEX2341. Bioresour. Technol. 175, 537–544.
Travieso, L., Canizares, R.O., Borja, R., Benitez, F., Dominguez, A.R., Dupeyron, R., Yen, H.-W., Chen, P.-W., Hsu, C.-Y., Lee, L., 2017. The use of autotrophic Chlorella vul-
Valiente, V., 1999. Heavy metal removal by microalgae. Bull. Environ. Contam. garis in chromium (VI) reduction under different reduction conditions. J. Taiwan
Toxicol. 62 (2), 144–151. Inst. Chem. Eng. 74, 1–6.
Tuzen, M., Sarı, A., Mendil, D., Uluozlu, O.D., Soylak, M., Dogan, M., 2009. Zakhama, S., Dhaouadi, H., M’Henni, F., 2011. Nonlinear modelisation of heavy metal
Characterization of biosorption process of As(III) on green algae Ulothrix cylin- removal from aqueous solution using Ulva lactuca algae. Bioresour. Technol. 102 (2),
dricum. J. Hazard. Mater. 165 (1), 566–572. 786–796.
Upadhyay, A.K., Mandotra, S.K., Kumar, N., Singh, N.K., Singh, L., Rai, U.N., 2016. Zhang, X., Zhao, X., Wan, C., Chen, B., Bai, F., 2016. Efficient biosorption of cadmium by
Augmentation of arsenic enhances lipid yield and defense responses in alga the self-flocculating microalga Scenedesmus obliquus AS-6-1. Algal Res. 16, 427–433.
Nannochloropsis sp. Bioresour. Technol. 221, 430–437. Zhang, Z., Yan, K., Zhang, L., Wang, Q., Guo, R., Yan, Z., Chen, J., 2019. A novel cad-
USEPA, 2016. National Primary Drinking Water Regulations. in: EPA 816-F-09-004, (Ed.) mium-containing wastewater treatment method: Bio-immobilization by microalgae
U.S.E.P. Agency, United State Environmental Protection Agency. United State. cell and their mechanism. J. Hazard. Mater. 374, 420–427.
Valdez, C., Perengüez, Y., Mátyás, B., Guevara, M.F., 2018. Analysis of removal of cad- Zhao, Y., Song, X., Yu, L., Han, B., Li, T., Yu, X., 2019. Influence of cadmium stress on the
mium by action of immobilized Chlorella sp. micro-algae in alginate beads. lipid production and cadmium bioresorption by Monoraphidium sp. QLY-1. Energy
F1000Research 7 54-54. Convers. Manage. 188, 76–85.
Wang, N.-X., Huang, B., Xu, S., Wei, Z.-B., Miao, A.-J., Ji, R., Yang, L.-Y., 2014. Effects of Zheng, H., Guo, W., Li, S., Wu, Q., Yin, R., Feng, X., Du, J., Ren, N., Chang, J.-S., 2016.
nitrogen and phosphorus on arsenite accumulation, oxidation, and toxicity in Biosorption of cadmium by a lipid extraction residue of lipid-rich microalgae. RSC
Chlamydomonas reinhardtii. Aquat. Toxicol. 157, 167–174. Adv. 6 (24), 20051–20057.
Wang, N.-X., Li, Y., Deng, X.-H., Miao, A.-J., Ji, R., Yang, L.-Y., 2013. Toxicity and Zhu, N., Zhang, J., Tang, J., Zhu, Y., Wu, Y., 2018. Arsenic removal by periphytic biofilm
bioaccumulation kinetics of arsenate in two freshwater green algae under different and its application combined with biochar. Bioresour. Technol. 248, 49–55.

11

You might also like