You are on page 1of 10

Bioresource Technology 246 (2017) 2–11

Contents lists available at ScienceDirect

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

Review

Recent developments on algal biochar production and characterization MARK


a a b c a
Kai Ling Yu , Beng Fye Lau , Pau Loke Show , Hwai Chyuan Ong , Tau Chuan Ling ,

Wei-Hsin Chend, Eng Poh Nge, Jo-Shu Changf,g,
a
Institute of Biological Sciences, Faculty of Science, University of Malaya, 50603 Kuala Lumpur, Malaysia
b
Bioseparation Research Group, Department of Chemical and Environmental Engineering, Faculty of Engineering, University of Nottingham Malaysia Campus, Jalan
Broga, 43500 Semenyih, Selangor Darul Ehsan, Malaysia
c
Department of Mechanical Engineering, Faculty of Engineering, University of Malaya, 50603 Kuala Lumpur, Malaysia
d
Department of Aeronautics and Astronautics, National Cheng Kung University, Tainan 701, Taiwan
e
School of Chemical Sciences, Universiti Sains Malaysia, USM, Malaysia
f
Department of Chemical Engineering, National Cheng Kung University, Tainan 701, Taiwan
g
Research Center for Energy Technology and Strategy, National Cheng Kung University, Tainan 701, Taiwan

G RA P H I C A L AB S T R A C T

A R T I C L E I N F O A B S T R A C T

Keywords: Algal biomass is known as a promising sustainable feedstock for the production of biofuels and other valuable
Algal biochar products. However, since last decade, massive amount of interests have turned to converting algal biomass into
Biochar synthesis biochar. Due to their high nutrient content and ion-exchange capacity, algal biochars can be used as soil
Biochar characterization amendment for agriculture purposes or adsorbents in wastewater treatment for the removal of organic or in-
Pyrolysis and torrefaction
organic pollutants. This review describes the conventional (e.g., slow and microwave-assisted pyrolysis) and
Thermochemical conversion
newly developed (e.g., hydrothermal carbonization and torrefaction) methods used for the synthesis of algae-
based biochars. The characterization of algal biochar and a comparison between algal biochar with biochar
produced from other feedstocks are also presented. This review aims to provide updated information on the
development of algal biochar in terms of the production methods and the characterization of its physical and
chemical properties to justify and to expand their potential applications.

1. Introduction to its wide distribution, rapid growth and high CO2 fixation efficiency
(Chen et al., 2015b; Roberts et al., 2015a). Growing algal biomass is a
Renewable energy such as biomass is receiving more attention on its distinctive way to sequester atmospheric carbon dioxide to minimize
applicability for sustainable supply to the environment (Al-Hamamre greenhouse effect (Gronnow et al., 2013). A potential improvement on
et al., 2017). Algal biomass is considered as a promising feedstock due conversion of biomass to biochar through thermochemical processes


Corresponding author at: Department of Chemical Engineering, National Cheng Kung University, Tainan 701, Taiwan.
E-mail address: changjs@mail.ncku.edu.tw (J.-S. Chang).

http://dx.doi.org/10.1016/j.biortech.2017.08.009
Received 11 June 2017; Received in revised form 2 August 2017; Accepted 3 August 2017
Available online 05 August 2017
0960-8524/ © 2017 Elsevier Ltd. All rights reserved.
K.L. Yu et al. Bioresource Technology 246 (2017) 2–11

such as pyrolysis, hydrothermal carbonization or torrefaction is com- the future without the dependence on fossil fuels and their growth can
mendable for further utilizing the algal biomass in the context of efficiently reduce emissions of greenhouse gases. Microalgae are pop-
biorefinery. Upgrading of current thermochemical processes is im- ular choice for biofuel production since it is easy to cultivate them in a
portant to enhance the economic feasibility of biochar production large amount under various environments (Vassilev and Vassileva,
process with a faster rate, higher yield, and better quality for applica- 2016). All types of biofuels (i.e., solid, liquid and gas biofuels) can be
tions (e.g., higher value applications of biochars other than as fertilizer, generated from microalgae using several conversion methods such as
such as adsorbent for environmental pollutant uptake or bio-materials direct combustion, chemical conversion, biochemical conversion and
for medical applications). thermochemical conversion. In addition to the success of algal tech-
Biochar is a carbon-rich product of biomass produced by thermal nology in fuel production, other value-added co-products such as bio-
decomposition under limited oxygen (O2) supply at a relatively low char can be produced simultaneously from algal feedstock for a bior-
temperature (Alhashimi and Aktas, 2017; Chang et al., 2015). The efinery concept (Foley et al., 2011; Rashid et al., 2013).
conversion of wet algal biomass into algal char can occur at a relatively Biochar is gaining more attention on its long term advantage in
moderate thermal condition in a short time under batch processing carbon sequestration and application in agriculture for soil amendment
conditions (Heilmann et al., 2010). The composition of biochar varies (Ennis et al., 2012). Biochar technology involved the CO2 uptake
with the types of feedstock. For example, microalgal biochar consists of through photosynthesis. The captured carbon undergoes conversion
large aggregates with irregular porosity. These features vary from the processes such as pyrolysis to produce biochar with characteristic of
structure of lignocellulose biochar produced after thermal decomposi- long-term carbon storage through soil amendment (Sohi et al., 2009).
tion treatment (Torri et al., 2011). Algal biochar has lower surface area Biochar production differs from other biomass energy production sys-
and carbon content, but higher cation exchange capacity compared to tems as this technology is carbon-negative. The International Biochar
lignocellulose biochar. The higher pH properties of algal biochar can Initiative (2008) estimated that biochar production has the potential of
balance acidified soils, while the higher nutrient content of nitrogen, mitigation of climate change by providing 3.67 Gt CO2 per year using
ash and inorganic elements are beneficial for soil amendment in agri- only biomass wastes. Biochar is potential to sequester up to 12% of
culture (Kołtowski et al., 2017; Sun et al., 2017). greenhouse gases from anthropogenic sources in ecologically and eco-
In addition to soil amendment utilization, biochar can also be used nomically sustainable systems (Ennis et al., 2012). Implementation of
as a bio-adsorbent in water treatment to remediate organic/inorganic biochar’s ability in mitigation of climate change at global scale is re-
contaminants due to its abundance of organic functional groups and cognized (Molina et al., 2009).
inorganic minerals (Awad et al., 2017; Johansson et al., 2016). How- As a strategy to store captured carbon for a long time on impact to
ever, selection of cheap raw bio-materials with better properties and greenhouse gas accumulation, biomass is converted into biochar that
easy preparation steps of bio-adsorbent still present major challenges has more than 90% carbon (Heilmann et al., 2010). Biochar derived
for the application of biochar in the conventional wastewater treatment from microalgae is nutrient-rich (especially nitrogen-rich) so is well-
(Inyang et al., 2016). Microalgae have been studied extensively as a suited to serve as a fertilizer in agriculture soil (Torri et al., 2011). With
potential biosorbent for the removal of organic and inorganic pollutants all the advantages of algal biochar, converting algal biomass into bio-
due to their biosorption capacity attributed by a large amount of char can be economically feasible for algal production enterprise
functional groups (Guo et al., 2016; Zeraatkar et al., 2016). Those (Bryant et al., 2012). However, to date there is still very limited lit-
specific functional groups on microalgal biochar lead to some special erature regarding algal biochar and its utilization. This may markedly
physic-chemical properties that could enhance its adsorption efficiency hinder the future development and application of algae-based biochar
for organic contaminants (Zheng et al., 2017). The biorefinery model of (Shukla et al., 2017), suggesting the need for a comprehensive litera-
microalgae that can produce valuable biofuels and biochar simulta- ture review in this promising area.
neously making it an interesting approach for future research and de-
velopment. Therefore, this review aims to provide the latest develop- 3. Algal biochar production
ment of algae-based biochar production technology using
thermochemical processes such as pyrolysis, torrefaction and hydro- Algal biomass is converted into biochar mainly through thermo-
thermal carbonization. This review also aims to provide updated in- chemical conversions, such as pyrolysis, hydrothermal carbonization
formation on the characterization of physical and chemical properties and torrefaction. The conventional way of synthesizing biochar is
of microalgal biochar to justify and to expand their potential applica- through slow pyrolysis which gives a high char yield. Hydrothermal
tions. A future recommendation of research work related to algal bio- carbonization produced a final carbonaceous material, which is also
char is also proposed in the review. denoted as hydrochar. Torrefaction is a pre-treatment method that has
been studied extensively to upgrade biomass into carbon-rich solid fuel
2. Algal biomass as the source for producing biofuels and biochar such as biochar (Chen et al., 2015d; Kumar et al., 2017). Details of
biochar synthesis methods are described as follows.
Fig. 1 shows the schematic view of algal biomass production in
renewable energy and carbon sequestration. With the ability in nu- 3.1. Conventional and microwave-assisted pyrolysis
trients uptake, algae possess a high growth rate and environmentally
tolerant characteristics to rapidly dominate in high nutrients environ- Pyrolysis is one of the most promising technologies for converting
ment. The high nutrient content of algae makes it a suitable feedstock biomass into valuable biofuels as well as biochars. There are several
for biochar production for the potential use in soil amendment and types of pyrolysis based on their operating conditions such as the
implement for long-term carbon sequestration. Algal biochar derived conventional slow pyrolysis, fast pyrolysis and also the latest developed
from the remediation of wastewater could provide a notable benefit in microwave-assisted pyrolysis. Each type of pyrolysis produces different
the future by utilizing biomass for carbon negative energy generation composition of solid, liquid and gaseous products according to their
and application to the environment (Bird et al., 2011). However, the operating parameters, such as temperature, heating rate and residence
high nutrient content of algal biomass makes it a disadvantage in pyr- time. Slow pyrolysis gives the highest yield of biochar while fast pyr-
olysis product distribution where more bio-oil product can be obtained. olysis gives the highest yield of bio-oil with biochar as by-product (Roy
One way to solve this problem is to extract the lipids from algae for bio- and Dias, 2017).
oil production, while the residues are further used for biochar pro-
duction to the context of biorefinery (Wang et al., 2013). Algae have 3.1.1. Conventional pyrolysis
been known as one of the promising sustainable energy feedstocks for Pyrolysis is the heating of biomass with the absence of oxygen or air

3
K.L. Yu et al. Bioresource Technology 246 (2017) 2–11

Fig. 1. Algal biomass production for renewable energy production and carbon sequestration.

at a given rate typically at a temperature range of around 300–700 °C formation of biochar product. Conventional pyrolysis used typical
(Chen et al., 2015c). The products obtained from pyrolysis are de- heating process, while microwave pyrolysis required some pre-treat-
termined by several factors, in particular the temperature and heating ment and catalysts prior to heating. Moreover, in conventional pyr-
rate (Basu, 2010). Biochar yield increases with a decrease in pyrolysis olysis, the pyrolysis gas is the by-product, while bio-oil, hot water and
temperature, an increase in the residence time, and a preferable low non-condensable gases can be obtained after condensation of micro-
heating rate. In addition, feedstock properties, such as moisture content wave pyrolysis. There are numerous studies on the pyrolysis of lig-
and particle size, also significantly affect the yield of biochar produced nocellulosic biomass but reports on the production of algal biochar via
via pyrolysis (Tripathi et al., 2016). Slow pyrolysis, the conventional microwave heating are limited (Wan et al., 2009). Biochar that un-
process in charcoal production, could yield the maximum amount of dergoes further chemical or thermal processing after production can be
biochar from biomass compared to other processes, such as fast pyr- transformed into activated carbon (Spokas et al., 2011). However,
olysis and gasification (Chaiwong et al., 2012; Mohan et al., 2014). Up previous study mentioned that the decrease of functional groups in
to 50% of the carbon from biomass may be stored in the stable biochar biochar due to the release of volatiles during pyrolysis would result in a
through pyrolysis (Bird et al., 2011). Biomass undergoes slow pyrolysis challenge when using it as an adsorbent (Wang et al., 2015). Micro-
process with a vapour residence time from several minutes to hours for wave-assisted pyrolysis can be used in future scale-up production from
char production (Du, 2013). Vapours are restrained and reacted with algal products into biochar for applications such as soil fertilizer due to
solid phase extensively for more char yield at the end of the process its economic production process. Previous reports on the algal biochar
(Mohan et al., 2014). Slow pyrolysis, in general, carried out at low production via the pyrolysis process are summarized in Table 1.
heating rates of 0.1–1 K/s with a residence time of around 450–550 s.
Pre-pyrolysis happens at the beginning, followed by solid decomposi-
tion corresponding to the high rate pyrolysis process to form pyrolysis 3.2. Torrefaction
products. Decomposition of the char finally occurs at a very low rate
and carbon-rich biochar is formed (Suganya et al., 2016). Most of the Torrefaction is a thermochemical conversion process which is per-
traditional slow pyrolysis used fixed bed reactors where heating is formed under atmospheric pressure at the temperature between 200
provided by heated surface but there are studies that looked into al- and 300 °C under an inert condition in the absence of oxygen (Chen
ternative heating methods such as microwave heating (Du, 2013; Wan et al., 2014a; Chen et al., 2015a). The process partly decomposes bio-
et al., 2009). mass and produces a solid product (called torrefied biomass or char)
with high carbon content. A general sketch of torrefaction process in
algal biochar production is shown in Fig. 2. Torrefaction is used for
3.1.2. Microwave-assisted pyrolysis biofuel production from microalgae and its prime purpose is to produce
Microwave-assisted pyrolysis is one of the most efficient thermo- biochar (Bach et al., 2017b). Torrefaction is an emerging thermal bio-
chemical processes in the production of biochar, bio-oil and syngas and mass pre-treatment process able to remove volatiles through different
it has been successfully applied to plant residues such as wood and decomposition reactions to reduce major limitations of biomass, up-
sewage sludge (Lei et al., 2011). Some of the advantages of microwave grade biomass quality and alter the combustion behaviour (Nhuchhen
pyrolysis are high products yield, reduction of harmful chemical in bio- et al., 2014). By altering the combustion behavior, fuel flexibility is
oil, energy and cost-saving. Microwave technology uses electro- enhanced by making a wide range of fuels efficiently applicable in a co-
magnetic waves to cause oscillation of material molecules and to pro- firing power plant. Torrefaction of microalgal biomass grown by using
duce heat. The technical advantages of microwave-assisted pyrolysis flue gas from the thermal power plant can be made suitable for co-firing
over conventional pyrolysis are (1) uniform microwave heating that is in a pulverized coal power (Wu et al., 2012). The thermal pre-treatment
applicable on larger biomass particles, (2) production of higher heating of torrefaction can be divided into dry and wet torrefaction, which is
value syngas that can be used for in-situ electricity for microwave also known as hydrothermal torrefaction or hydrothermal carboniza-
generation, (3) cleaner products due to no agitation and fluidization in tion (Chen et al., 2015d; Yan et al., 2009). Hydrothermal carbonization
the process, and (4) microwave heating is a mature technology with is to be discussed in Section 3.3. Comparison of the characteristics of
scale-up feasibility (Du, 2013). There are some differences between dry and wet torrefaction is shown in Table 2. The major advantage of
conventional slow pyrolysis and microwave-assisted pyrolysis in the wet torrefaction over dry torrefaction is its ability to produce energy-

4
K.L. Yu et al. Bioresource Technology 246 (2017) 2–11

Table 1
The pyrolysis processes used for algal biochar production.

Types of process Biomass feedstock Temperature (°C) Biochar production References

Slow pyrolysis Chlorella-based algal residue 300–700 • 56.3% at 300 °C Chang et al. (2015)
• 66.2% at 500 °C
• 65.0% at 700 °C
• High concentration of nitrogen and other inorganic
elements
Slow pyrolysis Brown Laminaria japonica macroalgae 200–800 • 78.34% at 200 °C Jung et al. (2016)
• 63.64% at 400 °C
• 37.96% at 600 °C
• 27.05% at 800 °C
Fixed-bed pyrolysis Chlorella vulgaris 300–900 • 19.3–43.46% of biochar yield on different temperatures Yuan et al. (2015)
Fixed-bed pyrolysis Chlamydomonas reinhardtii 350 • Nitrogen-rich biochar Torri et al. (2011)
• Largest fraction in term of mass, 44 ± 1% w/w mass yield
of biochar
Fixed-bed pyrolysis Scenedesmus dimorphus 300–600 • when
Surface area of biochar increased from 1.72 to 123 m /g
temperature increased from 300-500 °C; reduced to
2
Bordoloi et al. (2016)

89 m2/g at 600 °C
• The recalcitrance of biochar increased from 0.62 to 0.76
with increasing temperature
Fixed-bed pyrolysis Seaweed 250–600 • Biochar
250 °C
energy yield of 61.50% at 600 °C to 93.95% at Tag et al. (2016)

Fixed-bed pyrolysis Green macroalgae Cladophora glomerata 400–600 • 4440 wt% yield at 400 °C Norouzi et al. (2016)
• 39 wt% yield at 500 °C
• 31%wt%of biochar
yield at 600 °C
Fluidized-bed fast Defatted Chlorella vulgaris 500 • Energy recoveryyield Wang et al. (2013)
pyrolysis • 94% of algal biomass in bio-oil and biochar is

• High inorganic biochar content


Fluidized-bed pyrolysis Laminaria digitata, Fucus serratus and mix 500 • 29–36% of biochar yield Yanik et al. (2013)
macroalgae species from Black sea • Production of biochar with lower heating value due to
higher ash content of 41–52%
Stepwise and non-
stepwise pyrolysis
Wastewater treatment high rate algal
pond (WWT HRAP) biomass
300–500 • High amount of biochar with more than 50 wt% of the
initial biomass was produced under both heating regimes
Mehrabadi et al.
(2016)
Microwave pyrolysis Macroalgae 240–400 • 54.8% of biochar yield Shuttleworth et al.
(2012)

dense product within a short residence time due to high heat transfer Table 2
rate in the aqueous media (Coronella et al., 2012; Hoekman et al., Comparison of the characteristics of dry and wet torrefaction (Yan et al., 2009).
2013). In wet torrefaction, microalgae are treated under hot com-
Characteristics Torrefaction
pressed water, producing a solid product that has high calorific value,
better hydrophobicity, and lower ash content (Bach et al., 2017b). Dry Wet (Hydrothermal)
Torrefaction could produce biochar with high calorific value or higher
Temperature 200–300 °C 180–260 °C
heating value so it can be used as an alternative feedstock for clean
Media Inert nitrogen gas Hot compressed water
energy production other than fossil fuel. The hydrophobicity of biochar Pressure Atmospheric pressure 200–700 psi
originating from the surface functional groups and the lower ash con- Residence time 80 min 5 min
tent are important properties, determining its effectiveness in applying Cooling process Flowing nitrogen; indirect Immerse into ice bath
as an adsorbent for pollutants uptake from soil and water. Bach et al. water cooling rapidly
Additional processes – Filtration and evaporation
(2017a) showed that, after wet torrefaction, at least 61.5% of energy in
Energy density Lower Higher
the microalgal biomass is retained. The calorific value intensified up to
21% and there is a decrease in the ash content of the microalgae.

Fig. 2. General torrefaction process for algal biochar production.

5
K.L. Yu et al. Bioresource Technology 246 (2017) 2–11

Torrefaction is usually carried out at a low temperature and short re- where biochar is produced at lower temperatures (180–260 °C) and
sidence time under low heating rate to give a higher yield of solid elevated water or steam pressures (Libra et al., 2011; Titirici et al.,
product (Deng et al., 2009; Nhuchhen et al., 2014). Wu et al. (2012) 2012). The process takes place in water under a self-generated pressures
reported that the solid yield decreased when the torrefaction tem- being less than 10 bar with water as solvent (Titirici et al., 2012). This
perature is increased. The effect of residence time on the mass yield of process can be suitable for concentrating carbon of wet biomass where
torrefied biomass at 300 °C shows the mass yield decreased with an no drying is required prior to pyrolysis, making it a potential alternative
increase in the residence time. The study concluded that temperature for treatment of some waste streams (Brownsort, 2009). Char produced
influenced the mass yield more than residence time. Chen et al. (2014b) from hydrothermal carbonization is called hydrochar (Libra et al.,
shows the isothermal and non-isothermal torrefaction characteristics 2011). HTC produce a higher product in a shorter period of time and
and kinetics of a microalga Scenedesmus obliquus CNW-N. Microalgae requires lower energy expense than conventional carbonization process
are classified based on the torrefaction temperature, light, mild and (Tekin et al., 2014). The advantages of the HTC process include (1)
severe torrefaction from the maximum decomposition rate and weight required only low carbonization temperature, (2) can be synthesized in
loss. Non-isothermal torrefaction required intense pre-treatment than the aqueous phase, (3) inexpensive process, (4) renewable materials
the isothermal torrefaction. Pre-treatment severity is intensified by the can be used as sources such as biomass and for the use of value-added
increasing of heating rate in non-isothermal torrefaction. Uemura et al. chemicals, such as nanoparticles in the structure (Kubo, 2013). Char
(2015) reported the yields of solid, liquid and gas for a series of tor- product obtained from HTC has the following properties: uniform
refaction temperature on a macroalga Laminaria japonica. The solid spherical particles; controlled porosity; functional surfaces (e.g., eOH,
yield decreased when the torrefaction temperature was increased. The eC]O, eCOOH); easily controlled surface chemistry and electronic
decrease in solid yield may be attributed to the decomposition of two properties (Titirici et al., 2012).
major components, alginate and mannitol in L. japonica. However, both In a hydrothermal process, biomass can be converted into valuable
the liquid and gas yields increased when the temperature was increased products such as biochar, bio-oil and gaseous products by manipulating
in conjunction with a decrease of solid yield with torrefaction tem- process variables such as temperature, time of reaction, feedstock, the
perature. (Bach et al., 2017a) mentioned that the solid yield decreased presence of catalysts and pressure (Tekin et al., 2014). Temperature is
with an increase in temperature and residence time. The solid yield the most influential variable in the HTC process followed by residence
decreased from 61.68% to 52.58% when the temperature was increased time and the types of feedstock (Nizamuddin et al., 2017). Lower
from 160 °C to 180 °C. The solid yield decreased from 62.92% to temperature tends to give a higher yield of solid product compared to
51.84% when the residence time increased from 5 min to 30 min. Chen higher temperatures by affecting its physical and chemical character-
et al. (2016) showed that the solid yield of 51.3–93.9% in the torrefied istics. At a higher temperature, the carbon content is higher whereas the
microalgae residue at the temperature ranged from 200 to 300 °C with a hydrogen and oxygen contents are lower. This results in a formation of
residence time of 15–60 min. Previous study also shows the solid yield biochar with greater higher heating value (HHV). Char produced from
of 50.8–95.7% in microalgae Chlamydomonas sp. JSC4 residue after HTC of microalgae has a unique composition and with bituminous coal
torrefaction at temperature 200–300 °C for 15–60 min. It is re- quality (Heilmann et al., 2010). Process conditions were under a lower
commended that torrefaction of microalgae residue should be carried temperature of 200 °C with 0.5 h of reaction time for effective carbo-
out at an optimum temperature of 250 °C or below for less weight loss nization and production of algal char. The brief reaction time in the
and higher energy densification (Chen, 2015). The impact of torrefac- batch process suggested the development of a continuous process for
tion upon biomass properties has been extensively investigated in the HTC processing of algae. There are no specific catalytic agents that
last decade. However, there is limited literature on the study of algal significantly enhanced the carbonization process and/or increase the
biochar from torrefaction process. As torrified algal biomass is a high yield of biochar. The most conceivable alternative pathway proposed in
quality and environmental friendly solid product that may offer con- the study was carbonization via a dehydration route. Previous literature
siderable opportunities for worldwide greenhouse gas mitigation, fu- of algal hydrochar production in HTC is shown in Table 3. As there is
ture research on this area is suggested. very limited literature on algal char production from the HTC process, it
would be an interesting topic for future studies. HTC process offers the
advantages of lowering the production cost and shortening the time
3.3. Hydrothermal carbonization needed for the production of biochar. This can be achieved by utilizing
algal biomass residue and converting it into more valuable biofuels and
Hydrothermal carbonization (HTC) is a new thermochemical tech- other products. HTC represents a feasible alternative way to convert
nique that has gained more attention in the recent years due to its wet biomass into biochar by omitting the drying process.
environmental friendly and cost effectiveness (Erlach et al., 2012; Xiao
et al., 2012). HTC is a distinctive process that involves the conversion of
carbohydrate components of biomass into carbon-rich solids in water

Table 3
Algal hydrochar production using hydrothermal carbonization (HTC).

Biomass feedstock Temperature (°C) Reaction time Solid mass yield (%) References

Microalgae
Dunaliella salina 190–210 30–120 min 25.3–45.7 Heilmann et al. (2010)
Arthrospira platensis 190–210 2–4 h 21.6–36.7 Yao et al. (2016)
Nannochloropsis sp. 180–220 15–30 min 30–47 Lu et al. (2014)
Nannochloropsis oculata 180–215 15–45 min 41–51 Levine et al. (2013)
Whole Spirulina 175–215 30 min 23.3–49.3 Broch et al. (2014)
Lipid extracted Spirulina 175 30 min 44.6 Broch et al. (2014)

Macroalgae
Sargassum horneri 180–210 2–16 h 32.7–52.3 Xu et al. (2013)
Laminaria digitata 200–250 1h 18.4–21.8 Smith and Ross (2016)
Laminaria hyperborean 200–250 1h 23.6–39.0 Smith and Ross (2016)
Alaria esculenta 200–250 1h 23.7–30.0 Smith and Ross (2016)

6
K.L. Yu et al. Bioresource Technology 246 (2017) 2–11

Table 4
Physical properties of biochars derived from microalgal and macroalgal biomass.

Feedstock Pyrolysis temperature and Biochar yield (%) HHV (MJ/kg) BET surface area References
duration (m2/g)

Microalgae
Arthrospira platensis 550 °C, 60 min 31.0 15.8 n.d. Chaiwong et al. (2012)
Arthrospira platensis 500 °C, 20 min 25.0 n.d. n.d. Conti et al. (2016)
Chlamydomonas reinhardtii 350 °C, 20 min 44.0 13.0 n.d. Torri et al. (2011)
Chlorella vulgaris 350 °C 31.0 23.0 n.d. Wang et al. (2013)
Cladophora sp. n.a. n.d. n.d. n.d. Maddi et al. (2011)
Desmodesmus communis 500 °C, 20 min 34.0 n.d. n.d. Conti et al. (2016)
Isochrysis sp. 500 °C, 1 h ∼30 7.56 n.d. Aysu et al. (2016)
Lyngbya sp. n.a. ∼20 16.4 n.d. Maddi et al. (2011)
Nannochloropsis sp. 400, 500, and 600 °C, 60 min 24.8–33.5 14.8 n.d. Aysu and Sanna (2015)
Spirogyra sp. 550 °C, 60 min 28.0 23.0 n.d. Chaiwong et al. (2012)
Tetraselmis sp. 500 °C, 1 h ∼20 10.37 n.d. Aysu et al. (2016)
Lacustrine alga 400, 500 and 600 °C, 45 min 41.3–48.3 n.d. n.d. Chiodo et al. (2016)
Mixed consortia 500, 600, and 800 °C, 1, 2, and 3 h 33.0–63.0 n.d. n.d. Sarkar et al. (2015)

Macroalgae
Ascophyllum nodosum Microwave assisted thermal 21.4 21.2 n.d. Yuan and Macquarrie
treatment (2015)
Cladophora sp. 550 °C, 60 min 31.0 16.7 n.d. Chaiwong et al. (2012)
Cladophora vagabunda 250–400 °C 67.0 n.d. 8.3 Bird et al. (2012)
Cladophora coelothrix Hydrothermal liquefaction 10.4 18.3 n.d. Neveux et al. (2014)
Cladophora vagabunda Hydrothermal liquefaction 18.7 15.1 n.d. Neveux et al. (2014)
Chaetomorpha linum Hydrothermal liquefaction 8.4 20.5 n.d. Neveux et al. (2014)
Derbesia tenuissima Hydrothermal liquefaction 8.1 9.1 n.d. Neveux et al. (2014)
Eucheuma sp. 450 °C, 60 min 57.2–61.7 14.6–17.2 30.0–34.8 Roberts et al. (2015b)
Gracilaria sp. 450 °C, 60 min 59.8–61.8 11.1–16.1 2.0–3.6 Roberts et al. (2015b)
Gracilaria gracilis 400, 500, and 600 °C n.d. 11.4 n.d. Francavilla et al. (2015)
Gracilaria gracilis residue 400, 500, and 600 °C 26–32 14.5 n.d. Francavilla et al. (2015)
Kappaphycus sp. 450 °C, 60 min 54.1–59.2 13.0–17.8 2.2–2.8 Roberts et al. (2015b)
Oedogonium sp. Hydrothermal liquefaction 10.2 5.2 n.d. Neveux et al. (2014)
Porphyra tenera 500 °C, 60 min n.d. n.d. n.d. Park et al. (2016)
Saccharina sp. 450 °C, 60 min 45.3–49.7 11.4–14.8 1.3–8.5 Roberts et al. (2015b)
Sargassum sp. 450 °C, 60 min 49.0–61.9 11.8–13.5 2.5–7.5 Roberts et al. (2015b)
Ulva flexuousa 300–500 °C, 60 min n.d. n.d. 1.2–4.3 Bird et al. (2011)
Ulva ohnoi Hydrothermal liquefaction 12.1 4.0 n.d. Neveux et al. (2014)
Undaria sp. 450 °C, 60 min 60.3–62.4 10.7–14.7 1.3–8.9 Roberts et al. (2015b)
Mixture of Cladophora coelothrix, Chaetomorpha 250–400 °C 74.5 n.d. n.d. Bird et al. (2012)
indica, Ulva flexuosa
Kelp n.a. n.d. n.d. 26.6 Boakye et al. (2016)
Spray-dried algae 300, 450, 600, and 750 °C, 10 and 19.3–72.8 8.2–9.2 14–19 Ronsse et al. (2013)
60 min

n.a., not available; n.d., not determined.

4. Algal biochar characterization composition of cell wall. Besides that, the high ash content in algal
samples might act catalytically during pyrolysis to alter the distribution
Characterization of the physical and chemical properties of algal of products.
biochar is of great importance in determining their potential applica-
tions. The physical and chemical properties of biochar produced from 4.1.2. Higher heating value
both microalgae and macroalgae are discussed as follows. The higher heating value (HHV) is the amount of heat released
when the sample (fuel) is combusted and the products have returned to
4.1. Physical properties of algal biochar a temperature of 25 °C. It is typically measured using a bomb calori-
meter in the laboratory. As shown in Table 4, biochars derived from
4.1.1. Biochar yield microalgae (7.6–23.0 MJ/kg) and macroalgae (5.2–21.2 MJ/kg) have
Previous studies have demonstrated that algal samples could yield comparable HHV. In general, the HHV of algal biochars is lower than
relatively high amounts of biochar per unit biomass. However, com- that of lignocellulosic biochars (which has been reported to exceed
parative analysis of biochar derived from microalgae and macroalgae in 30 MJ/kg), is likely attributed to a lower carbon content but higher ash
terms of their yields under similar experimental conditions is still content in algal biomass.
lacking. As shown in Table 4, the yield of biochar derived from micro-
and macroalgae (on a dry weight basis) ranged from 20.0 to 63.0% and 4.1.3. Surface properties
8.1–62.4%, respectively. There are attempts to investigate the factors The physical properties of biochars, such as their surface area,
which affect the yield of algal biochar. Ronsse et al. (2013), for in- porosity, and pore volume, are routinely analyzed. The Brunauer-
stance, reported that the yield of algal biochar is lower than lig- Emmet-Teller (BET) surface area of biochar can be estimated using a
nocellulosic biomass, namely pine wood, wheat straw, and green waste. surface area analyzer while a scanning electron microscope (SEM) can
The yield decreased when the pyrolysis temperature and residence time be used to profile the surface topography and particle structure. The
were increased from 300 to 750 °C and from 10 to 60 min, respectively. surface area of algal biochar is generally low but some studies reported
Similar results were also observed by Tag et al. (2016) who postulated that an increase in the pyrolysis temperature could result in a higher
that such observations could be due to the differences in the chemical surface area (Bird et al., 2011). Ronsse et al. (2013) found that the ash
components between algal and terrestrial biomass, particularly for the content in the biomass is negatively correlated with the surface area of

7
K.L. Yu et al. Bioresource Technology 246 (2017) 2–11

the resulting biochar. Wang et al. (2013) reported that the surface area study by Tag et al. (2016), algal biochar produced at different pyrolysis
of biochar derived from C. vulgaris (2.4 m2/g) was deemed to be low temperature was found to have higher CEC (25.6–52.6 cmol/kg) than
relative to biochar obtained from lignocellulosic biomass. Working with that produced from vine pruning (32.2–61.0 cmol/kg) and orange po-
biochar derived from macroalgae, Roberts et al. (2015b) reported that mace (25.6–52.6 cmol/kg) under similar experimental conditions. As
biochar produced from the macroalgae Eucheuma sp. has a significantly CEC correlated with the ash content, it was suggested that the alkali and
higher surface area (30.03–34.82 m2/g) than those of other species earth alkali metals in biomass promoted the formation of O-containing
which ranged from 1.29 to 8.87 m2/g. surface functional groups in the resulting biochar (Cely et al., 2015).
Findings thus far have indicated that biochar derived from algal
samples has distinct structure and characteristics when compared with
the biomass before subjected to pyrolysis. For instance, Wang et al. 4.2.5. Functional groups
(2013) observed that the particles of biochar derived from a green Fourier transform infrared spectroscopy (FTIR) spectral analysis is
microalga, Chlorella vulgaris, are compact and irregular, and these are used to identify the various functional groups in algal biochar. This
different from the structure of the feedstock before pyrolysis. Similar technique can be used to compare the chemical profiles of biochar
results were also reported for the biochar derived from Chlamydomonas produced under different conditions, as well as between the raw bio-
reinhardtii (Torri et al. (2011). These observations are in contrast with mass and the resulting biochar. The latter is exemplified by the work of
biochars from lignocellulosic biomass, which normally retain their Biswas et al. (2017) whereby FT-IR spectra demonstrated that, fol-
feedstock’s structure. Table 4 summarizes the physical properties of lowing hydrothermal liquefaction, the macromolecular crystal structure
biochars derived from several microalgae and macroalgae species. of the algae was disrupted and products (biochar and bio-oil) were
formed. Table 5 summarizes the chemical properties of biochar derived
4.2. Chemical properties of algal biochar from some microalgae and macroalgae species.

4.2.1. Proximate and ultimate analysis


In proximate analysis, the moisture, volatile matter, ash, and fixed 4.3. Factors affecting the properties of algal biochar
carbon content of the samples are measured. Ultimate analysis (also
termed as elemental analysis), on the other hand, includes the de- In general, the method and temperature of pyrolysis are two im-
termination of carbon, hydrogen, nitrogen, sulphur, and oxygen within portant factors that determine the physical and chemical properties of
the samples. As shown in Table 5, algal biochars are typically low in C biochar (Jindo et al., 2014; Mukome et al., 2013). The operating con-
but high in N and minerals (ash) contents. The low C content, in ditions such as heating rate, reaction vessel, chemical activation, re-
comparison to those derived from lignocellulosic biomass, is the char- sidence time, and highest treatment temperature (HTT) also influence
acteristic of biochars produced from micro- and macroalgae. the properties of biochar produced. The HHT is regarded to have the
greatest effect on the physical properties of biochar produced (Mukome
4.2.2. Inorganic elemental analysis et al., 2013). In a recent study, Palanisamy et al. (2017) reported that
Inorganic mineral content of algal biochar is typically determined biochar of C. vulgaris prepared at higher temperatures (450–600 °C)
using the inductively coupled plasma atomic emission spectrometry contained a higher proportion of organic matter (C, H and N) than those
(ICP-AES). Previous work has provided insights into the abundance of produced at lower temperatures.
various inorganic elements in algal biochar and enables comparison In addition, the type of feedstock used may affect the chemical
with biochar derived from other feedstock. For example, Wang et al. composition of the resulting biochar, such as the ash content, pH, H/C
(2013) reported that biochar derived from C. vulgaris contained higher ratio, surface area, as well as cation and anion exchange capabilities
concentrations of various trace elements, including P, K, Mg and Ca, (Sun et al., 2014; Tag et al., 2016). These factors are of great im-
compared to biochar from lignocellulosic biomass. In a separate study, portance when comes to the applications of biochar, particularly in the
biochars produced from both red and brown seaweeds have high con- agriculture sector. In a comparative analysis in the properties of biochar
centrations of N (0.3–2.8%), P (0.5–6.6 g/kg), and K (5.1–119 g/kg) produced by algae, grass, manure, wood, pomace, and nutshell, it was
(Roberts et al., 2015b). Results from both studies, when taken together, found that feedstock is a better predictor of variation in the ash content
seem to suggest that algal biochar has good prospects as fertilizers in and C/N ratio of biochar than pyrolysis temperature (Mukome et al.,
view of their mineral contents. 2013). However, when one feedstock is being considered, pyrolysis
temperature is the best predictor for the surface area.
4.2.3. pH In terms of biochar production from algal species, the proportion
As shown in Table 5, all algal biochar was found to be alkaline. The and composition of nutrients of algal-derived biochar is likely influ-
pH of biochar derived from macroalgae ranges from 7.6 to 13.7, enced by a number of abiotic and biotic factors like species, habitat (e.g.
whereas little information is available regarding those derived from fresh, brackish or saline environment), and other factors. It is evidenced
microalgae. The pH of biochar is found to be affected by the pyrolysis by the findings of a number of researchers. Dealing with seaweed bio-
temperature and the algal samples. According to the study by Ronsse chars, Roberts et al. (2015b) found that biochars produced from red and
et al. (2013), the pH of biochar is correlated to the presence of oxygen brown seaweeds have different elemental composition with biochars
functionalities in the biochar. Tag et al. (2016) noted that the pH of from red seaweeds have higher concentrations of S and K and lower
algal biochar (8.7–13.7) increased with an increase in the pyrolysis concentrations of C and H than biochars produced from the brown
temperature (250–600 °C). A plausible explanation is the increase in the seaweeds. There are also variations in most physicochemical char-
relative ash content in the biochar caused a rise in the pH of the biochar acteristics of the biochar between species collected from different lo-
especially under severe pyrolysis conditions (Ronsse et al., 2013) cations especially the cation exchange capability of the biochars. For
instance, the exchangeable K in Undaria sp. (13–420 cmol/kg) and
4.2.4. Cation exchange capacity Kappaphycus sp. (26–210 cmol/kg) differed by an order of magnitude
The cation exchange capacity (CEC) of biochar is a measure of the depending on the location. Bird et al. (2012) also noted a difference in
ability of biochar to adsorb cation nutrients. In other words, a biochar the yield and ash content of the Chadophora vagabunda with their earlier
with high CEC exerts beneficial effect by preventing nutrient leaching work (Bird et al., 2011). These may be attributed to moving to pilot
in the soil. Roberts et al. (2015b) showed that seaweed biochar samples scale production with commercial pyrolysis equipment rather than
have negligible or no exchangeable Al, but have high levels of the re- controlled laboratory production, purity of the materials collected in
maining exchangeable cations such as Ca, K, Mg and Na. In another the field, and lesser control over pyrolysis conditions.

8
K.L. Yu et al. Bioresource Technology 246 (2017) 2–11

Table 5
Chemical properties of biochars derived from microalgal and macroalgal biomass.

Biochar Proximate analysis (wt%) Ultimate analysis (wt%, dry ash-free) pH References

Moisture Volatile Ash Fixed carbon C H N S O


matter

Microalgae
Arthrospira platensis – 7.6 47.8 44.6 45.3 1.2 2.6 0.1 0.3 – Chaiwong et al. (2012)
Arthrospira platensis – – 47.8 – 45.3 1.2 2.6 – – – Chaiwong et al. (2013)
Arthrospira platensis – 28.0 29.5 42.0 51.0 2.5 7.7 – 18.0 – Conti et al. (2016)
Cladophora sp. – – – – 62.7 2.2 4.0 – 31.0 – Maddi et al. (2011)
Chlamydomonas – – 45.0 – 40.0 1.4 5.3 < 0.1 9.3 – Torri et al. (2011)
reinhardtii
Chlorella vulgaris 3.4 23.5 20.0 54.2 62.0 3.9 9.4 – 4.8 – Wang et al. (2013)
Desmodesmus communis – 22.0 36.8 41.0 51.0 3.0 7.0 0.1 14.0 – Conti et al. (2016)
Isochrysis sp. – – – – 43.5 1.5 3.7 – 51.3 – Aysu et al. (2016)
Lyngbya sp. – – – – 68.8 2.5 6.7 – 21.9 – Maddi et al. (2011)
Nannochloropsis sp. – – – – 55.3 2.0 5.2 – 37.5 – Aysu and Sanna (2015)
Spirogyra sp. – 16.8 23.5 59.7 62.4 0.4 2.1 0.5 4.1 – Chaiwong et al. (2012)
Tetraselmis sp. – – – – 48.1 1.7 4.3 – 46.0 – Aysu et al. (2016)
Lacustrine algae – – – – 27.0–30.5 0.8–1.1 0.7–1.3 – 67.1–71.5 10.05 Chiodo et al. (2016)

Macroalgae
Ascophyllum nodosum – – – – 51.1 5.3 1.9 – – – Yuan and Macquarrie
(2015)
Chaetomorpha indica – – 73.5 – 10.2 0.8 1.1 – – 7.8 Bird et al. (2011)
Chaetomorpha linum – – 16.0 – 23.6 1.3 2.4 – – 9.6 Bird et al. (2011)
Chaetomorpha linum – – – – 48.1 5.3 2.7 0.8 25.5 – Neveux et al. (2014)
Cladophora sp. – – – – 51.1 0.6 2.0 1.9 0.7 – Chaiwong et al. (2012)
Cladophora coelothrix – – – – 34.6 1.5 3.3 – – 8.7 Bird et al. (2011)
Cladophora coelothrix – – – – 44.2 4.1 3.7 1.7 20.1 – Neveux et al. (2014)
Cladophora patentiramea – – 47.0 – 20.3 1.2 1.7 – – 9.1 Bird et al. (2011)
Cladophora vagabunda – – 54.2 – 21.8 1.2 2.0 – – 9.9 Bird et al. (2011)
Cladophora vagabunda – – – – 36.1 3.9 2.6 0.3 20.6 – Neveux et al. (2014)
Cladophora patentiramea – – 47.0 – 20.3 1.2 1.7 – – 9.1 Bird et al. (2011)
Cladophoropsis sp. – – 46.5 – 23.6 1.5 2.8 – – 10.1 Bird et al. (2011)
Caulerpa taxifolia – – 20.9 – 24.8 1.2 2.4 – – 9.7 Bird et al. (2011)
Derbesia tenuissima – – – – 19.9 3.0 1.3 9.3 22.1 – Neveux et al. (2014)
Eucheuma sp. – – – – 23.7–25.6 1.2–1.8 0.7–0.8 7.0–9.3 20.6–24.9 8.2–8.6 Roberts et al. (2015b)
Gracilaria sp. – – – – 24.5–30.9 1.5–2.2 1.3–2.8 2.7–4.4 16.5–19.8 7.6–8.1 Roberts et al. (2015b)
Gracilaria gracilis 4.1 – 47.4 – 36.8 0.2 2.3 2.6 10.6 – Francavilla et al. (2015)
Gracilaria gracilis residue 5.4 – 30.7 – 41.2 2.2 2.3 1.3 22.3 – Francavilla et al. (2015)
Kappaphycus sp. – – – – 22.2–31.3 1.1–2.1 0.3–0.7 5.5–6.8 15.6–23.8 8.8–9.0 Roberts et al. (2015b)
Oedogonium sp. – – – – 12.2 1.4 1.2 0 6.6 – Neveux et al. (2014)
Porphyra tenera 5.5 – 21.3 73.2 74.7 3.0 9.1 4.2 9.0 – Park et al. (2016)
Saccharina sp. – – – – 28.0–35.0 1.9–2.4 2.2–2.4 1.0–1.6 16.4–18.4 11.0–11.2 Roberts et al. (2015b)
Sargassum sp. – – – – 28.9–29.1 2.0–2.1 1.0–1.1 0.9–2.8 15.3–18.2 10.1–10.8 Roberts et al. (2015b)
Ulva flexuousa – – 26.5–42.6 – 22.3–30.9 1.1–3.2 4.0–7.7 – – 8.0–10.1 Bird et al. (2011)
Ulva ohnoi – – – – 9.6 2.3 0.9 16.3 35.1 – Neveux et al. (2014)
Undaria sp. – – – – 27.3–34.8 1.7–2.8 2.3–2.4 0.6–0.8 14.1–15.6 9.9–10.9 Roberts et al. (2015b)
Mixture – 11.6–54.2 22.9–42.7 22.9–45.7 36.6–45.3 1.4–3.7 1.4–2.1 0–0.8 9.3–34.7 8.7–13.7 Tag et al. (2016)
Kelp – – – – 27.4 0.6 1.28 – 12.1 – Boakye et al. (2016)
Spray-dried algae – 3.9–70.0 46.3–76.4 30.0–96.1 62.7–90.6 1.4–7.2 – – – 4.9–12.5 Ronsse et al. (2013)

–: not determined.

4.4. Comparison of algal biochar with biochars derived from different lignocellulosic biochar gives important implications in their applica-
feedstocks tions. The algal biochar is likely to provide significant direct nutrient
benefits to soils and crop productivity, and are likely to be particularly
It is generally accepted that the first critical step in determining the useful for application on acidic soils. However, they are volumetrically
utility and applications of algal biochar is the quantification of the less able to provide the carbon sequestration benefits that can be pro-
properties of algal biochar and a comparison with biochar produced vided with the high-carbon-content lignocellulosic biochars. Therefore,
from other commonly used terrestrial biomass feedstocks (Bird et al., Roberts et al. (2015b) proposed that a blending of seaweed and lig-
2011). Studies thus far have pointed to the fact that biochars produced nocellulosic biochars could provide a soil ameliorant that combines a
from algal samples are fundamentally different from those produced high fixed C content with a mineral-rich substrate to enhance crop
from lignocellulosic feedstock (Bird et al., 2012; Maddi et al., 2011). In productivity.
general, biochars produced from various algal species tend to have low
carbon content, surface area and cation exchange capability but high in
5. Future challenges and opportunities
pH, nitrogen, and extractable inorganic nutrients including P, K, Ca and
Mg. In contrast, lignocellulosic materials-derived biochars tend to have
Algal biochar could influence the global carbon sequestration on
higher carbon contents and cation exchange capabilities with pH values
mitigation of climate change from the production of energy to the ap-
usually lower than 7, and significantly lower ash and available nutrient
plications in environmental management. In particular, the use of algal
contents (Jindo et al., 2014).
biochar as biosorbents in wastewater treatment should be studied fur-
The difference in the chemical profiles between algal and other
ther. Nevertheless, it is important to take note of some of the challenges

9
K.L. Yu et al. Bioresource Technology 246 (2017) 2–11

that may lie ahead, such as cost effectiveness of the process and the fact Bach, Q.-V., Chen, W.-H., Lin, S.-C., Sheen, H.-K., Chang, J.-S., 2017a. Wet torrefaction of
microalga Chlorella vulgaris ESP-31 with microwave-assisted heating. Energy
that wastewater contains various impurities that may interfere with the Convers. Manage. 141, 163–170.
process. In addition, production of algal biochar via the existing Bach, Q.-V., Chen, W.-H., Sheen, H.-K., Chang, J.-S., 2017b. Gasification kinetics of raw
methods could be further enhanced with microwave-assisted tech- and wet-torrefied microalgae Chlorella vulgaris ESP-31 in carbon dioxide. Bioresour.
Technol.
nology, whereas the additional energy consumption associated with the Basu, P., 2010. Chapter 3 – Pyrolysis and torrefaction. In: Biomass Gasification and
microwave-assisted process should be well considered and minimized. Pyrolysis. Academic Press, Boston, pp. 65–96.
Microwave-assisted technology in the production of algal biochar is Bird, M.I., Wurster, C.M., de Paula Silva, P.H., Bass, A.M., de Nys, R., 2011. Algal biochar
– production and properties. Bioresour. Technol. 102 (2), 1886–1891.
indeed a promising approach that deserves more attention in the future. Bird, M.I., Wurster, C.M., De Paula Silva, P.H., Paul, N.A., De Nys, R., 2012. Algal biochar:
Furthermore, optimization of algal biochar productivity can also be effects and applications. GCB Bioenergy 4 (1), 61–69.
achieved by employing a continuous process. Biswas, B., Arun Kumar, A., Bisht, Y., Singh, R., Kumar, J., Bhaskar, T., 2017. Effects of
temperature and solvent on hydrothermal liquefaction of Sargassum tenerrimum
In terms of the algal biomass to be used for biochar production,
algae. Bioresour. Technol.
microalgae seem to be a more promising feedstock when compared to Boakye, P., Lee, C.W., Lee, W.M., Woo, S.H., 2016. The cell viability on kelp and fir
macroalgae, in view of their advantages over the latter, such as their biochar and the effect on the field cultivation of corn. Clean Technol. 22 (1), 29–34.
rapid growth and the ease of cultivation and harvesting. Macroalgae, on Bordoloi, N., Narzari, R., Sut, D., Saikia, R., Chutia, R.S., Kataki, R., 2016.
Characterization of bio-oil and its sub-fractions from pyrolysis of Scenedesmus di-
the other hand, require a larger area for cultivation and generally re- morphus. Renewable Energy.
quire a longer period of cultivation time. The quality and quantity of Broch, A., Jena, U., Hoekman, S., Langford, J., 2014. Analysis of solid and aqueous phase
macroalgal biomass are also greatly dependent on varieties of abiotic products from hydrothermal carbonization of whole and lipid-extracted algae.
Energies 7 (1), 62.
and biotic factors. Moreover, microalgae are deemed to have higher Brownsort, P.A., 2009. Biomass Pyrolysis Processes: Review of Scope, Control and
potential in biochar production due to their high nutrient contents Variability. UK Biochar Research Center, Edinburgh.
which are suitable for the use in agriculture. The recycling of microalgal Bryant, H.L., Gogichaishvili, I., Anderson, D., Richardson, J.W., Sawyer, J., Wickersham,
T., Drewery, M.L., 2012. The value of post-extracted algae residue. Algal Res. 1 (2),
residue after biofuel production for a biorefinery concept is also fea- 185–193.
sible. Taken together, future research on different aspects pertaining to Cely, P., Gascó, G., Paz-Ferreiro, J., Méndez, A., 2015. Agronomic properties of biochars
microalgal biochar, such as production methods and newer applica- from different manure wastes. J. Anal. Appl. Pyrol. 111, 173–182.
Chaiwong, K., Kiatsiriroat, T., Vorayos, N., Thararax, C., 2012. Biochar production from
tions, is highly recommended in view of the scarcity of information on
freshwater algae by slow pyrolysis. Maejo Int. J. Sci. Technol. 6 (2).
these aspects. Chaiwong, K., Kiatsiriroat, T., Vorayos, N., Thararax, C., 2013. Study of bio-oil and bio-
char production from algae by slow pyrolysis. Biomass Bioenergy 56, 600–606.
Chang, Y.-M., Tsai, W.-T., Li, M.-H., 2015. Chemical characterization of char derived from
6. Conclusions
slow pyrolysis of microalgal residue. J. Anal. Appl. Pyrol. 111, 88–93.
Chen, W.-H., 2015. Microalgae oil: algae cultivation and harvest, algae residue torre-
Algal biomass as a third generation feedstock can be utilized in faction and diesel engine emissions tests. Aerosol Air Qual. Res.
biochar production in the context of biorefinery. Pyrolysis, torrefaction Chen, W.-H., Huang, M.-Y., Chang, J.-S., Chen, C.-Y., 2014a. Thermal decomposition
dynamics and severity of microalgae residues in torrefaction. Bioresour. Technol.
or hydrothermal carbonization are suitable processes to produce algal 169, 258–264.
biochar of different composition and properties for varying further Chen, W.-H., Huang, M.-Y., Chang, J.-S., Chen, C.-Y., 2015a. Torrefaction operation and
applications. Microwave-assisted technology could be an interesting optimization of microalga residue for energy densification and utilization. Appl.
Energy 154, 622–630.
approach in enhancing the current biochar production technologies. Chen, W.-H., Huang, M.-Y., Chang, J.-S., Chen, C.-Y., Lee, W.-J., 2015b. An energy ana-
Characterization of algal biochar is important for the understanding of lysis of torrefaction for upgrading microalga residue as a solid fuel. Bioresour.
their chemical and physical properties, which are useful for de- Technol. 185, 285–293.
Chen, W.-H., Lin, B.-J., Huang, M.-Y., Chang, J.-S., 2015c. Thermochemical conversion of
termining their potential applications (such as fertilizer for agricultural microalgal biomass into biofuels: a review. Bioresour. Technol. 184, 314–327.
purpose or adsorbent for water treatment). Development of algal bio- Chen, W.-H., Peng, J., Bi, X.T., 2015d. A state-of-the-art review of biomass torrefaction,
char technologies is expected to contribute to a more sustainable en- densification and applications. Renew. Sustain. Energy Rev. 44, 847–866.
Chen, W.-H., Wu, Z.-Y., Chang, J.-S., 2014b. Isothermal and non-isothermal torrefaction
vironment in the future. characteristics and kinetics of microalga Scenedesmus obliquus CNW-N. Bioresour.
Technol. 155, 245–251.
Acknowledgements Chen, Y.-C., Chen, W.-H., Lin, B.-J., Chang, J.-S., Ong, H.C., 2016. Impact of torrefaction
on the composition, structure and reactivity of a microalga residue. Appl. Energy 181,
110–119.
This study is supported by University of Malaya under the SATU Chiodo, V., Zafarana, G., Maisano, S., Freni, S., Urbani, F., 2016. Pyrolysis of different
Joint Research Scheme (RU018J-2016, RU018L-2016, RU018O-2016 biomass: direct comparison among Posidonia Oceanica, Lacustrine Alga and White-
,RU018C-2016, ST001-2017, ST002-2017, ST003-2017, ST004-2017, Pine. Fuel 164, 220–227.
Conti, R., Fabbri, D., Vassura, I., Ferroni, L., 2016. Comparison of chemical and physical
ST005-2017, ST006-2017 and RP031B-15AET). The financial support indices of thermal stability of biochars from different biomass by analytical pyrolysis
from Taiwan’s Ministry of Science and Technology (MOST) under grant and thermogravimetry. J. Anal. Appl. Pyrol. 122, 160–168.
numbers of MOST 106-3113-E-006-011, 106-3113-E-006-004-CC2, Coronella, C.J., Yan, W., Reza, M.T., Vasquez, V.R., 2012. Method for Wet Torrefaction of
a Biomass. Google Patents.
104-2221-E-006-227-MY3, and 103-2221-E-006-190-MY3 is also Deng, J., Wang, G.-J., Kuang, J.-H., Zhang, Y.-L., Luo, Y.-H., 2009. Pretreatment of
greatly appreciated. agricultural residues for co-gasification via torrefaction. J. Anal. Appl. Pyrol. 86 (2),
331–337.
Du, Z., 2013. Thermochemical Conversion of Microalgae for Biofuel Production.
References University of Minnesota.
Ennis, C.J., Evans, A.G., Islam, M., Ralebitso-Senior, T.K., Senior, E., 2012. Biochar:
Al-Hamamre, Z., Saidan, M., Hararah, M., Rawajfeh, K., Alkhasawneh, H.E., Al-Shannag, carbon sequestration, land remediation, and impacts on soil microbiology. Crit. Rev.
M., 2017. Wastes and biomass materials as sustainable-renewable energy resources Environ. Sci. Technol. 42 (22), 2311–2364.
for Jordan. Renew. Sustain. Energy Rev. 67, 295–314. Erlach, B., Harder, B., Tsatsaronis, G., 2012. Combined hydrothermal carbonization and
Alhashimi, H.A., Aktas, C.B., 2017. Life cycle environmental and economic performance gasification of biomass with carbon capture. Energy 45 (1), 329–338.
of biochar compared with activated carbon: a meta-analysis. Resour. Conserv. Recycl. Foley, P.M., Beach, E.S., Zimmerman, J.B., 2011. Algae as a source of renewable che-
118, 13–26. micals: opportunities and challenges. Green Chem. 13 (6), 1399–1405.
Awad, Y.M., Lee, S.-E., Ahmed, M.B.M., Vu, N.T., Farooq, M., Kim, I.S., Kim, H.S., Francavilla, M., Manara, P., Kamaterou, P., Monteleone, M., Zabaniotou, A., 2015.
Vithanage, M., Usman, A.R.A., Al-Wabel, M., Meers, E., Kwon, E.E., Ok, Y.S., 2017. Cascade approach of red macroalgae Gracilaria gracilis sustainable valorization by
Biochar, a potential hydroponic growth substrate, enhances the nutritional status and extraction of phycobiliproteins and pyrolysis of residue. Bioresour. Technol. 184,
growth of leafy vegetables. J. Cleaner Prod. 156, 581–588. 305–313.
Aysu, T., Abd Rahman, N.A., Sanna, A., 2016. Catalytic pyrolysis of Tetraselmis and Gronnow, M.J., Budarin, V.L., Mašek, O., Crombie, K.N., Brownsort, P.A., Shuttleworth,
Isochrysis microalgae by nickel ceria based catalysts for hydrocarbon production. P.S., Hurst, P.R., Clark, J.H., 2013. Torrefaction/biochar production by microwave
Energy 103, 205–214. and conventional slow pyrolysis–comparison of energy properties. GCB Bioenergy 5
Aysu, T., Sanna, A., 2015. Nannochloropsis algae pyrolysis with ceria-based catalysts for (2), 144–152.
production of high-quality bio-oils. Bioresour. Technol. 194, 108–116. Guo, W.-Q., Zheng, H.-S., Li, S., Du, J.-S., Feng, X.-C., Yin, R.-L., Wu, Q.-L., Ren, N.-Q.,

10
K.L. Yu et al. Bioresource Technology 246 (2017) 2–11

Chang, J.-S., 2016. Removal of cephalosporin antibiotics 7-ACA from wastewater of slow pyrolysis biochar: influence of feedstock type and pyrolysis conditions. GCB
during the cultivation of lipid-accumulating microalgae. Bioresour. Technol. 221, Bioenergy 5 (2), 104–115.
284–290. Roy, P., Dias, G., 2017. Prospects for pyrolysis technologies in the bioenergy sector: a
Heilmann, S.M., Davis, H.T., Jader, L.R., Lefebvre, P.A., Sadowsky, M.J., Schendel, F.J., review. Renew. Sustain. Energy Rev. 77, 59–69.
Von Keitz, M.G., Valentas, K.J., 2010. Hydrothermal carbonization of microalgae. Sarkar, O., Agarwal, M., Naresh Kumar, A., Venkata Mohan, S., 2015. Retrofitting he-
Biomass Bioenergy 34 (6), 875–882. trotrophically cultivated algae biomass as pyrolytic feedstock for biogas, bio-char and
Hoekman, S.K., Broch, A., Robbins, C., Zielinska, B., Felix, L., 2013. Hydrothermal car- bio-oil production encompassing biorefinery. Bioresour. Technol. 178, 132–138.
bonization (HTC) of selected woody and herbaceous biomass feedstocks. Biomass Shukla, S., Gita, S., Bharti, V., Bhuvaneswari, G., Wikramasinghe, W., 2017. Atmospheric
Convers. Biorefin. 3 (2), 113–126. Carbon Sequestration Through Microalgae: Status, Prospects, and Challenges. In:
Inyang, M.I., Gao, B., Yao, Y., Xue, Y., Zimmerman, A., Mosa, A., Pullammanappallil, P., Agro-Environmental Sustainability. Springer, pp. 219–235.
Ok, Y.S., Cao, X., 2016. A review of biochar as a low-cost adsorbent for aqueous Shuttleworth, P., Budarin, V., Gronnow, M., Clark, J.H., Luque, R., 2012. Low tempera-
heavy metal removal. Crit. Rev. Environ. Sci. Technol. 46 (4), 406–433. ture microwave-assisted vs conventional pyrolysis of various biomass feedstocks. J.
Jindo, K., Mizumoto, H., Sawada, Y., Sanchez-Monedero, M.A., Sonoki, T., 2014. Physical Nat. Gas Chem. 21 (3), 270–274.
and chemical characterization of biochars derived from different agricultural re- Smith, A.M., Ross, A.B., 2016. Production of bio-coal, bio-methane and fertilizer from
sidues. Biogeosciences 11 (23), 6613–6621. seaweed via hydrothermal carbonisation. Algal Res. 16, 1–11.
Johansson, C.L., Paul, N.A., de Nys, R., Roberts, D.A., 2016. Simultaneous biosorption of Sohi, S., Loez-Capel, E., Krull, E., Bol, R., 2009. Biochar’s roles in soil and climate change:
selenium, arsenic and molybdenum with modified algal-based biochars. J. Environ. a review of research needs. CSIRO Land Water Sci. Rep. 5 (09), 1–57.
Manage. 165, 117–123. Spokas, K.A., Novak, J.M., Stewart, C.E., Cantrell, K.B., Uchimiya, M., DuSaire, M.G., Ro,
Jung, K.-W., Jeong, T.-U., Kang, H.-J., Ahn, K.-H., 2016. Characteristics of biochar de- K.S., 2011. Qualitative analysis of volatile organic compounds on biochar.
rived from marine macroalgae and fabrication of granular biochar by entrapment in Chemosphere 85 (5), 869–882.
calcium-alginate beads for phosphate removal from aqueous solution. Bioresour. Suganya, T., Varman, M., Masjuki, H., Renganathan, S., 2016. Macroalgae and microalgae
Technol. 211, 108–116. as a potential source for commercial applications along with biofuels production: a
Kołtowski, M., Charmas, B., Skubiszewska-Zięba, J., Oleszczuk, P., 2017. Effect of biochar biorefinery approach. Renew. Sustain. Energy Rev. 55, 909–941.
activation by different methods on toxicity of soil contaminated by industrial activity. Sun, H., Lu, H., Chu, L., Shao, H., Shi, W., 2017. Biochar applied with appropriate rates
Ecotoxicol. Environ. Saf. 136, 119–125. can reduce N leaching, keep N retention and not increase NH3 volatilization in a
Kubo, S., 2013. Nanostructured carbohydrate–derived carbonaceous materials. TANSO coastal saline soil. Sci. Total Environ. 575, 820–825.
2013 (258), 232–233. Sun, Y., Gao, B., Yao, Y., Fang, J., Zhang, M., Zhou, Y., Chen, H., Yang, L., 2014. Effects of
Kumar, G., Shobana, S., Chen, W.-H., Bach, Q.-V., Kim, S.-H., Atabani, A.E., Chang, J.-S., feedstock type, production method, and pyrolysis temperature on biochar and hy-
2017. A review of thermochemical conversion of microalgal biomass for biofuels: drochar properties. Chem. Eng. J. 240, 574–578.
chemistry and processes. Green Chem. 19 (1), 44–67. Tag, A.T., Duman, G., Ucar, S., Yanik, J., 2016. Effects of feedstock type and pyrolysis
Lei, H., Ren, S., Wang, L., Bu, Q., Julson, J., Holladay, J., Ruan, R., 2011. Microwave temperature on potential applications of biochar. J. Anal. Appl. Pyrol. 120, 200–206.
pyrolysis of distillers dried grain with solubles (DDGS) for biofuel production. Tekin, K., Karagöz, S., Bektaş, S., 2014. A review of hydrothermal biomass processing.
Bioresour. Technol. 102 (10), 6208–6213. Renew. Sustain. Energy Rev. 40, 673–687.
Levine, R.B., Sierra, C.O.S., Hockstad, R., Obeid, W., Hatcher, P.G., Savage, P.E., 2013. Titirici, M.-M., White, R.J., Falco, C., Sevilla, M., 2012. Black perspectives for a green
The use of hydrothermal carbonization to recycle nutrients in algal biofuel produc- future: hydrothermal carbons for environment protection and energy storage. Energy
tion. Environ. Prog. Sustainable Energy 32 (4), 962–975. Environ. Sci. 5 (5), 6796–6822.
Libra, J.A., Ro, K.S., Kammann, C., Funke, A., Berge, N.D., Neubauer, Y., Titirici, M.-M., Torri, C., Samorì, C., Adamiano, A., Fabbri, D., Faraloni, C., Torzillo, G., 2011.
Fühner, C., Bens, O., Kern, J., 2011. Hydrothermal carbonization of biomass re- Preliminary investigation on the production of fuels and bio-char from
siduals: a comparative review of the chemistry, processes and applications of wet and Chlamydomonas reinhardtii biomass residue after bio-hydrogen production.
dry pyrolysis. Biofuels 2 (1), 71–106. Bioresour. Technol. 102 (18), 8707–8713.
Lu, Y., Levine, R.B., Savage, P.E., 2014. Fatty acids for nutraceuticals and biofuels from Tripathi, M., Sahu, J.N., Ganesan, P., 2016. Effect of process parameters on production of
hydrothermal carbonization of microalgae. Ind. Eng. Chem. Res. 54 (16), 4066–4071. biochar from biomass waste through pyrolysis: a review. Renew. Sustain. Energy Rev.
Maddi, B., Viamajala, S., Varanasi, S., 2011. Comparative study of pyrolysis of algal 55, 467–481.
biomass from natural lake blooms with lignocellulosic biomass. Bioresour. Technol. Uemura, Y., Matsumoto, R., Saadon, S., Matsumura, Y., 2015. A study on torrefaction of
102 (23), 11018–11026. Laminaria japonica. Fuel Process. Technol. 138, 133–138.
Mehrabadi, A., Craggs, R., Farid, M.M., 2016. Pyrolysis of wastewater treatment high rate Vassilev, S.V., Vassileva, C.G., 2016. Composition, properties and challenges of algae
algal pond (WWT HRAP) biomass. Algal Res. biomass for biofuel application: an overview. Fuel 181, 1–33.
Mohan, D., Sarswat, A., Ok, Y.S., Pittman, C.U., 2014. Organic and inorganic con- Wan, Y., Chen, P., Zhang, B., Yang, C., Liu, Y., Lin, X., Ruan, R., 2009. Microwave-assisted
taminants removal from water with biochar, a renewable, low cost and sustainable pyrolysis of biomass: catalysts to improve product selectivity. J. Anal. Appl. Pyrol. 86
adsorbent–a critical review. Bioresour. Technol. 160, 191–202. (1), 161–167.
Molina, M., Zaelke, D., Sarma, K.M., Andersen, S.O., Ramanathan, V., Kaniaru, D., 2009. Wang, K., Brown, R.C., Homsy, S., Martinez, L., Sidhu, S.S., 2013. Fast pyrolysis of mi-
Reducing abrupt climate change risk using the Montreal Protocol and other reg- croalgae remnants in a fluidized bed reactor for bio-oil and biochar production.
ulatory actions to complement cuts in CO2 emissions. Proc. Natl. Acad. Sci. U.S.A. Bioresour. Technol. 127, 494–499.
106 (49), 20616–20621. Wang, N., Tahmasebi, A., Yu, J., Xu, J., Huang, F., Mamaeva, A., 2015. A Comparative
Mukome, F.N., Zhang, X., Silva, L.C., Six, J., Parikh, S.J., 2013. Use of chemical and study of microwave-induced pyrolysis of lignocellulosic and algal biomass. Bioresour.
physical characteristics to investigate trends in biochar feedstocks. J. Agric. Food Technol. 190, 89–96.
Chem. 61 (9), 2196–2204. Wu, K.-T., Tsai, C.-J., Chen, C.-S., Chen, H.-W., 2012. The characteristics of torrefied
Neveux, N., Yuen, A.K., Jazrawi, C., Magnusson, M., Haynes, B.S., Masters, A.F., Montoya, microalgae. Appl. Energy 100, 52–57.
A., Paul, N.A., Maschmeyer, T., de Nys, R., 2014. Biocrude yield and productivity Xiao, L.-P., Shi, Z.-J., Xu, F., Sun, R.-C., 2012. Hydrothermal carbonization of lig-
from the hydrothermal liquefaction of marine and freshwater green macroalgae. nocellulosic biomass. Bioresour. Technol. 118, 619–623.
Bioresour. Technol. 155, 334–341. Xu, Q., Qian, Q., Quek, A., Ai, N., Zeng, G., Wang, J., 2013. Hydrothermal carbonization
Nhuchhen, D.R., Basu, P., Acharya, B., 2014. A comprehensive review on biomass tor- of macroalgae and the effects of experimental parameters on the properties of hy-
refaction. Int. J. Renew. Energy Biofuels 2014, 1–56. drochars. ACS Sustain. Chem. Eng. 1 (9), 1092–1101.
Nizamuddin, S., Baloch, H.A., Griffin, G., Mubarak, N., Bhutto, A.W., Abro, R., Mazari, Yan, W., Acharjee, T.C., Coronella, C.J., Vásquez, V.R., 2009. Thermal pretreatment of
S.A., Ali, B.S., 2017. An overview of effect of process parameters on hydrothermal lignocellulosic biomass. Environ. Prog. Sustain. Energy 28 (3), 435–440.
carbonization of biomass. Renew. Sustain. Energy Rev. 73, 1289–1299. Yanik, J., Stahl, R., Troeger, N., Sinag, A., 2013. Pyrolysis of algal biomass. J. Anal. Appl.
Norouzi, O., Jafarian, S., Safari, F., Tavasoli, A., Nejati, B., 2016. Promotion of hydrogen- Pyrol. 103, 134–141.
rich gas and phenolic-rich bio-oil production from green macroalgae Cladophora Yao, C., Wu, P., Pan, Y., Lu, H., Chi, L., Meng, Y., Cao, X., Xue, S., Yang, X., 2016.
glomerata via pyrolysis over its bio-char. Bioresour. Technol. 219, 643–651. Evaluation of the integrated hydrothermal carbonization-algal cultivation process for
Palanisamy, M., Mukund, S., Sivakumar, U., Sivasubramanian, Karthikeyan, V., 2017. enhanced nitrogen utilization in Arthrospira platensis production. Bioresour.
Bio-char production from micro algal biomass of Chlorella vulgaris. PHYKOS 47 (1), Technol. 216, 381–390.
99–104. Yuan, T., Tahmasebi, A., Yu, J., 2015. Comparative study on pyrolysis of lignocellulosic
Park, S.H., Cho, H.J., Ryu, C., Park, Y.-K., 2016. Removal of copper (II) in aqueous so- and algal biomass using a thermogravimetric and a fixed-bed reactor. Bioresour.
lution using pyrolytic biochars derived from red macroalga Porphyra tenera. J. Ind. Technol. 175, 333–341.
Eng. Chem. 36, 314–319. Yuan, Y., Macquarrie, D.J., 2015. Microwave assisted step-by-step process for the pro-
Rashid, N., Rehman, M.S.U., Han, J.-I., 2013. Recycling and reuse of spent microalgal duction of fucoidan, alginate sodium, sugars and biochar from Ascophyllum nodosum
biomass for sustainable biofuels. Biochem. Eng. J. 75, 101–107. through a biorefinery concept. Bioresour. Technol. 198, 819–827.
Roberts, D.A., Cole, A.J., Paul, N.A., de Nys, R., 2015a. Algal biochar enhances the re- Zeraatkar, A.K., Ahmadzadeh, H., Talebi, A.F., Moheimani, N.R., McHenry, M.P., 2016.
vegetation of stockpiled mine soils with native grass. J. Environ. Manage. 161, Potential use of algae for heavy metal bioremediation, a critical review. J. Environ.
173–180. Manag. 181, 817–831.
Roberts, D.A., Paul, N.A., Dworjanyn, S.A., Bird, M.I., de Nys, R., 2015b. Biochar from Zheng, H., Guo, W., Li, S., Chen, Y., Wu, Q., Feng, X., Yin, R., Ho, S.-H., Ren, N., Chang, J.-
commercially cultivated seaweed for soil amelioration. Sci. Rep. 5, 9665. S., 2017. Adsorption of p-nitrophenols (PNP) on microalgal biochar: analysis of high
Ronsse, F., van Hecke, S., Dickinson, D., Prins, W., 2013. Production and characterization adsorption capacity and mechanism. Bioresour. Technol.

11

You might also like