Accepted Manuscript: Bioresource Technology

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 43

Accepted Manuscript

Review

Recent developments on algal biochar production and characterization

Kai Ling Yu, Beng Fye Lau, Pau Loke Show, Hwai Chyuan Ong, Tau Chuan
Ling, Wei-Hsin Chen, Ng Eng Poh, Jo-Shu Chang

PII: S0960-8524(17)31314-7
DOI: http://dx.doi.org/10.1016/j.biortech.2017.08.009
Reference: BITE 18618

To appear in: Bioresource Technology

Received Date: 11 June 2017


Revised Date: 2 August 2017
Accepted Date: 3 August 2017

Please cite this article as: Yu, K.L., Lau, B.F., Show, P.L., Ong, H.C., Ling, T.C., Chen, W-H., Poh, N.E., Chang,
J-S., Recent developments on algal biochar production and characterization, Bioresource Technology (2017), doi:
http://dx.doi.org/10.1016/j.biortech.2017.08.009

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Recent developments on algal biochar production and characterization

Kai Ling Yua, Beng Fye Laua, Pau Loke Showb, Hwai Chyuan Ongc, Tau Chuan Linga, Wei-

Hsin Chend, Ng Eng Pohe, 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, Bioseparation Research Group, 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.

1
Abstract

Algal biomass is known as a promising sustainable feedstock for the production of

biofuels and other valuable products. However, since last decade, massive amount of interests

have turned to converting algal biomass into biochar. Due to their high nutrient content and

ion-exchange capacity, algal biochars can be used as soil amendment for agriculture purposes

or adsorbents in wastewater treatment for the removal of organic or inorganic pollutants. This

review describes the conventional (e.g., slow and microwave-assisted pyrolysis) and 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.

Keywords: Algal biochar; biochar synthesis; biochar characterization; pyrolysis and

torrefaction; thermochemical conversion

2
1. Introduction

Renewable energy such as biomass is receiving more attention on its applicability for

sustainable supply to the environment (Al-Hamamre et al., 2017). Algal biomass is

considered as a promising feedstock due to its wide distribution, rapid growth and high CO2

fixation efficiency (Chen et al., 2015b; Roberts et al., 2015a). Growing algal biomass is a

distinctive way to sequester atmospheric carbon dioxide to minimize greenhouse effect

(Gronnow et al., 2013). A potential improvement on conversion of biomass to biochar

through thermochemical processes such as pyrolysis, hydrothermal carbonization or

torrefaction is commendable for further utilizing the algal biomass in the context of

biorefinery. Upgrading of current thermochemical processes is important to enhance the

economic feasibility of biochar production process with a faster rate, higher yield, and better

quality for applications (e.g., higher value applications of biochars other than as fertilizer,

such as adsorbent for environmental pollutant uptake or bio-materials for medical

applications)

Biochar is a carbon-rich product of biomass produced by thermal decomposition

under limited oxygen (O2) supply at a relatively low temperature (Alhashimi & Aktas, 2017;

Chang et al., 2015). The conversion of wet algal biomass into algal char can occur at a

relatively moderate thermal condition in a short time under batch processing conditions

(Heilmann et al., 2010). The composition of biochar varies with the types of feedstock. For

example, microalgal biochar consists of large aggregates with irregular porosity. These

features vary from the structure of lignocellulose biochar produced after thermal

decomposition treatment (Torri et al., 2011). Algal biochar has lower surface area and carbon

content, but higher cation exchange capacity compared to lignocellulose biochar. The higher

pH properties of algal biochar can balance acidified soils, while the higher nutrient content of

3
nitrogen, ash and inorganic elements are beneficial for soil amendment in agriculture

(Kołtowski et al., 2017; Sun et al., 2017).

In addition to soil amendment utilization, biochar can also be used as a bio-adsorbent

in water treatment to remediate organic/inorganic contaminants due to its abundance of

organic functional groups and inorganic minerals (Awad et al., 2017; Johansson et al., 2016).

However, selection of cheap raw bio-materials with better properties and easy preparation

steps of bio-adsorbent still present major challenges for the application of biochar in the

conventional wastewater treatment (Inyang et al., 2016). Microalgae have been studied

extensively as a potential biosorbent for the removal of organic and inorganic pollutants due

to their biosorption capacity attributed by a large amount of functional groups (Guo et al.,

2016; Zeraatkar et al., 2016). Those specific functional groups on microalgal biochar lead to

some special physic-chemical properties that could enhance its adsorption efficiency for

organic contaminants (Zheng et al., 2017). The biorefinery model of microalgae that can

produce valuable biofuels and biochar simultaneously making it an interesting approach for

future research and development. Therefore, this review aims to provide the latest

development of algae-based biochar production technology using thermochemical processes

such as pyrolysis, torrefaction and hydrothermal carbonization. This review also aims to

provide updated information on the characterization of physical and chemical properties of

microalgal biochar to justify and to expand their potential applications. A future

recommendation of research work related to algal biochar is also proposed in the review.

2. Algal biomass as the source for producing biofuels and biochar

Figure 1 shows the schematic view of algal biomass production in renewable energy

and carbon sequestration. With the ability in nutrients uptake, algae possess a high growth

rate and environmentally tolerant characteristics to rapidly dominate in high nutrients

4
environment. The high nutrient content of algae makes it a suitable feedstock for biochar

production for the potential use in soil amendment and implement for long-term carbon

sequestration. Algal biochar derived from the remediation of wastewater could provide a

notable benefit in the future by utilizing biomass for carbon negative energy generation and

application to the environment (Bird et al., 2011). However, the high nutrient content of algal

biomass makes it a disadvantage in pyrolysis product distribution where more bio-oil product

can be obtained. One way to solve this problem is to extract the lipids from algae for bio-oil

production, while the residues are further used for biochar production to the context of

biorefinery (Wang et al., 2013). Algae have been known as one of the promising sustainable

energy feedstocks for the future without the dependence on fossil fuels and their growth can

efficiently reduce emissions of greenhouse gases. Microalgae are popular choice for biofuel

production since it is easy to cultivate them in a large amount under various environments

(Vassilev & Vassileva, 2016). All types of biofuels (i.e., solid, liquid and gas biofuels) can be

generated from microalgae using several conversion methods such as direct combustion,

chemical conversion, biochemical conversion and thermochemical conversion. In addition to

the success of algal technology in fuel production, other value-added co-products such as biochar

can be produced simultaneously from algal feedstock for a biorefinery concept (Foley et al., 2011);

(Rashid et al., 2013).

Biochar is gaining more attention on its long term advantage in carbon sequestration

and application in agriculture for soil amendment (Ennis et al., 2012). Biochar technology

involved the CO2 uptake through photosynthesis. The captured carbon undergoes conversion

processes such as pyrolysis to produce biochar with characteristic of long-term carbon

storage through soil amendment (Sohi et al., 2009). Biochar production differs from other

biomass energy production systems as this technology is carbon-negative. The International

Biochar Initiative (2008) estimated that biochar production has the potential of mitigation of

5
climate change by providing 3.67 Gt CO2 per year using only biomass wastes. Biochar is

potential to sequester up to 12% of greenhouse gases from anthropogenic sources in

ecologically and economically sustainable systems (Ennis et al., 2012). Implementation of

biochar’s ability in mitigation of climate change at global scale is recognized (Molina et al.,

2009).

As a strategy to store captured carbon for a long time on impact to greenhouse gas

accumulation, biomass is converted into biochar that has more than 90% carbon (Heilmann et

al., 2010). Biochar derived from microalgae is nutrient-rich (especially nitrogen-rich) so is

well-suited to serve as a fertilizer in agriculture soil (Torri et al., 2011). With all the

advantages of algal biochar, converting algal biomass into biochar can be economically

feasible for algal production enterprise (Bryant et al., 2012). However, to date there is still

very limited literature regarding algal biochar and its utilization. This may markedly hinder

the future development and application of algae-based biochar (Shukla et al., 2017),

suggesting the need for a comprehensive literature review in this promising area.

3. Algal biochar production

Algal biomass is converted into biochar mainly through thermochemical conversions,

such as pyrolysis, hydrothermal carbonization and torrefaction. The conventional way of

synthesizing biochar is through slow pyrolysis which gives a high char yield. Hydrothermal

carbonization produced a final carbonaceous material, which is also denoted as hydrochar.

Torrefaction is a pre-treatment method that has been studied extensively to upgrade biomass

into carbon-rich solid fuel such as biochar (Chen et al., 2015d; Kumar et al., 2017). Details of

biochar synthesis methods are described as follows.

3.1 Conventional and microwave-assisted pyrolysis

6
Pyrolysis is one of the most promising technologies for converting biomass into

valuable biofuels as well as biochars. There are several types of pyrolysis based on their

operating conditions such as the conventional slow pyrolysis, fast pyrolysis and also the latest

developed microwave-assisted pyrolysis. Each type of pyrolysis produces different

composition of solid, liquid and gaseous products according to their operating parameters,

such as temperature, heating rate and residence time. Slow pyrolysis gives the highest yield

of biochar while fast pyrolysis gives the highest yield of bio-oil with biochar as by-product

(Roy & Dias, 2017).

3.1.1 Conventional pyrolysis

Pyrolysis is the heating of biomass with the absence of oxygen or air at a given rate

typically at a temperature range of around 300-700 ℃ (Chen et al., 2015c). The products

obtained from pyrolysis are determined by several factors, in particular the temperature and

heating rate (Basu, 2010). Biochar yield increases with a decrease in pyrolysis temperature,

an increase in the residence time, and a preferable low heating rate. In addition, feedstock

properties, such as moisture content and particle size, also significantly affect the yield of

biochar produced via pyrolysis (Tripathi et al., 2016). Slow pyrolysis, the conventional

process in charcoal production, could yield the maximum amount of biochar from biomass

compared to other processes, such as fast pyrolysis and gasification (Chaiwong et al., 2012;

Mohan et al., 2014). Up to 50% of the carbon from biomass may be stored in the stable

biochar through pyrolysis (Bird et al., 2011). Biomass undergoes slow pyrolysis process with

a vapour residence time from several minutes to hours for char production (Du, 2013).

Vapours are restrained and reacted with solid phase extensively for more char yield at the end

of the process (Mohan et al., 2014). Slow pyrolysis is, in general, carried at low heating rates

of 0.1-1 K/s with a residence time of around 450-550 s. Pre-pyrolysis happens at the

7
beginning, followed by solid decomposition corresponding to the high rate pyrolysis process

to form pyrolysis 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 traditional slow pyrolysis

used fixed bed reactors where heating is provided by heated surface but there are studies that

looked into alternative heating methods such as microwave heating (Du, 2013; Wan et al.,

2009).

3.1.2 Microwave-assisted pyrolysis

Microwave-assisted pyrolysis is one of the most efficient thermochemical processes

in the production of biochar, bio-oil and syngas and it has been successfully applied to plant

residues, such as wood and sewage sludge (Lei et al., 2011). Some of the advantages of

microwave pyrolysis are high products yield, reduction of harmful chemical in bio-oil, energy

and cost-saving. Microwave technology uses electromagnetic waves to cause oscillation of

material molecules and to produce heat. The technical advantages of microwave-assisted

pyrolysis over conventional pyrolysis are (1) uniform microwave heating that is applicable on

larger biomass particles, (2) production of higher heating value syngas that can be used for

in-situ electricity for microwave generation, (3) cleaner products due to no agitation and

fluidization in the process, and (4) microwave heating is a mature technology with scale-up

feasibility (Du, 2013). There are some differences between conventional slow pyrolysis and

microwave-assisted pyrolysis in the formation of biochar product. Conventional pyrolysis

used typical heating process, while microwave pyrolysis required some pre-treatment and

catalysts prior to heating. Moreover, in conventional pyrolysis, the pyrolysis gas is the by-

product, while bio-oil, hot water and non-condensable gases can be obtained after

condensation of microwave pyrolysis. There are numerous studies on the pyrolysis of

lignocellulosic biomass but reports on the production of algal biochar via microwave heating

are limited (Wan et al., 2009). Biochar that undergoes further chemical or thermal processing

8
after production can be transformed into activated carbon (Spokas et al., 2011). However,

previous study mentioned that the decrease of functional groups in biochar due to the release

of volatiles during pyrolysis would result in a challenge when using it as an adsorbent (Wang

et al., 2015). Microwave-assisted pyrolysis can be used in future scale-up production from

algal products into biochar for applications such as soil fertilizer due to its economic

production process. Previous reports on the algal biochar production via the pyrolysis process

are summarized in Table 1.

3.2 Torrefaction

Torrefaction is a thermochemical conversion process which is performed under

atmospheric pressure at the temperature between 200 and 300℃ under an inert condition in

the absence of oxygen (Chen et al., 2014a; Chen et al., 2015a). The process partly

decomposes biomass 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 Figure 2. Torrefaction is used for biofuel production from microalgae and its prime

purpose is to produce biochar (Bach et al., 2017b). Torrefaction is an emerging thermal

biomass pre-treatment process able to remove volatiles through different decomposition

reactions to reduce major limitations of biomass, upgrade biomass quality and alter the

combustion behaviour (Nhuchhen et al., 2014). By altering the combustion behavior, fuel

flexibility is enhanced by making a wide range of fuels efficiently applicable in a co-firing

power plant. Torrefaction of microalgal biomass grown by using flue gas from the thermal

power plant can be made suitable for co-firing in a pulverized coal power (Wu et al., 2012).

The thermal pre-treatment of torrefaction can be divided into dry and wet torrefaction, which

is also known as hydrothermal torrefaction or hydrothermal carbonization (Chen et al., 2015d;

Yan et al., 2009). Hydrothermal carbonization is to be discussed in Section 3.3. Comparison

of the characteristics of dry and wet torrefaction is shown in Table 2. The major advantage of

9
wet torrefaction over dry torrefaction is its ability to produce energy-dense product within a

short residence time due to high heat transfer rate in the aqueous media (Coronella et al.,

2012; Hoekman et al., 2013). In wet torrefaction, microalgae are treated under hot

compressed water, producing a solid product that has high calorific value, better

hydrophobicity, and lower ash content (Bach et al., 2017b). Torrefaction could produce

biochar with high calorific value or higher heating value so it can be used as an alternative

feedstock for clean energy production than fossil fuel. The hydrophobicity of biochar

originating from the surface functional groups and the lower ash content are important

properties, determining its effectiveness in applying as an adsorbent for pollutants uptake

from soil and water. Bach et al. (2017a) showed that, after wet torrefaction, at least 61.5% of

energy in the microalgal biomass is retained. The calorific value intensified up to 21 % and

there is a decrease in the ash content of the microalgae. Torrefaction is usually carried out at a

low temperature and short residence time under low heating rate to give a higher yield of

solid product (Deng et al., 2009; Nhuchhen et al., 2014). (Wu et al., 2012) reported that the

solid yield decreased when the torrefaction temperature is increased. The effect of residence

time on the mass yield of torrefied biomass at 300 ℃ shows the mass yield decreased with an

increase in the residence time. The study concluded that temperature influenced the mass

yield more than residence time. (Chen et al., 2014b) shows the isothermal and non-isothermal

torrefaction characteristics and kinetics of a microalga Scenedesmus obliquus CNW-N.

Microalgae are classified based on the torrefaction temperature, light, mild and severe

torrefaction from the maximum decomposition rate and weight loss. Non-isothermal

torrefaction required intense pre-treatment than the isothermal torrefaction. Pre-treatment

severity is intensified by the increasing of heating rate in non-isothermal torrefaction.

Uemura et al. (2015) reported the yields of solid, liquid and gas for a series of torrefaction

temperature on a macroalga Laminaria japonica. The solid yield decreased when the

10
torrefaction temperature was increased. The decrease in solid yield may be attributed to the

decomposition of two major components, alginate and mannitol in L. japonica. However,

both the liquid and gas yields increased when the temperature was increased in conjunction

with a decrease of solid yield with torrefaction temperature. (Bach et al., 2017a) mentioned

that the solid yield decreased with an increase in temperature and residence time. The solid

yield decreased from 61.68% to 52.58% when the temperature was increased from 160 °C to

180 °C. The solid yield decreased from 62.92% to 51.84% when the residence time increased

from 5 min to 30 min. Chen et al. (2016) showed that the solid yield of 51.3-93.9% in the

torrefied microalgae residue at the temperature ranged from 200-300 ℃ with a residence time

of 15-60 min. Previous study also shows the solid yield of 50.8-95.7% in microalgae

Chlamydomonas sp. JSC4 residue after torrefaction at temperature 200-300 ℃ for 15-60 min.

It is recommended that torrefaction of microalgae residue should be carried out at an

optimum temperature of 250 ℃ or below for less weight loss and higher energy densification

(Chen, 2015). The impact of torrefaction upon biomass properties has been extensively

investigated in the last decade. However, there is limited literature on the study of algal

biochar from torrefaction process. As torrified algal biomass is a high quality and

environmental friendly solid product that may offer considerable opportunities for worldwide

greenhouse gas mitigation, future research on this area is suggested.

3.3 Hydrothermal carbonization

Hydrothermal carbonization (HTC) is a new thermochemical technique that has

gained more attention in the recent years due to its 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 where biochar is produced at lower temperatures (180-260 ℃) and elevated water or

steam pressures (Libra et al., 2011; Titirici et al., 2012). The process takes place in water

11
under a self-generated pressures being less than 10 bar with water as solvent (Titirici et al.,

2012). This process can be suitable for concentrating carbon of wet biomass where no drying

is required prior to pyrolysis, making it a potential alternative for treatment of some waste

streams (Brownsort, 2009). Char produced from hydrothermal carbonization is called

hydrochar (Libra et al., 2011). HTC produce a higher product in a shorter period of time and

requires lower energy expense than conventional carbonization process (Tekin et al., 2014).

The advantages of the HTC process include (1) required only low carbonization temperature,

(2) can be synthesized in the aqueous phase, (3) inexpensive process, (4) renewable materials

can be used as sources such as biomass and for the use of value-added chemicals, such as

nanoparticles in the structure (Kubo, 2013). Char product obtained from HTC has the

following properties: uniform spherical particles; controlled porosity; functional surfaces

(e.g., -OH, -C=O, -COOH); easily controlled surface chemistry and electronic properties

(Titirici et al., 2012).

In a hydrothermal process, biomass can be converted into valuable products such as

biochar, bio-oil and gaseous products by manipulating process variables such as temperature,

time of reaction, feedstock, the presence of catalysts and pressure (Tekin et al., 2014).

Temperature is the most influential variable in the HTC process followed by residence time

and types of feedstock (Nizamuddin et al., 2017). Lower temperature tends to give a higher

yield of solid product compared to higher temperatures by affecting its physical and chemical

characteristics. At a higher temperature, the carbon content is higher whereas the hydrogen

and oxygen content is lower. This results in a formation of biochar with greater higher

heating value (HHV). Char produced from HTC of microalgae has a unique composition and

with bituminous coal quality (Heilmann et al., 2010). Process conditions were under a lower

temperature of 200 ℃ with 0.5 h of reaction time for effective carbonization and production

of algal char. The brief reaction time in the batch process suggested the development of a

12
continuous process for HTC processing of algae. There are no specific catalytic agents that

significantly enhanced the carbonization process and/or increase the yield of biochar. The

most conceivable alternative pathway proposed in the study was carbonization via a

dehydration route. Previous literature of algal hydrochar production in HTC is shown in

Table 3. As there is 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 needed for the production of biochar.

This can be achieved by utilizing algal biomass residue and converting it into more valuable

biofuels and other products. HTC represents a feasible alternative way to convert wet

biomass into biochar by omitting the drying process.

4. Algal biochar characterization

Characterization of the physical and chemical properties of algal biochar is of great

importance in determining their potential applications. The physical and chemical properties

of biochar produced from both microalgae and macroalgae are discussed as follows.

4.1 Physical properties of algal biochar

4.1.1 Biochar yield

Previous studies have demonstrated that algal samples could yield relatively high

amounts of biochar per unit biomass. However, comparative analysis of biochar derived from

microalgae and macroalgae in terms of their yields under similar experimental conditions is

still lacking. As shown in Table 4, the yield of biochar derived from micro- and macroalgae

(on a dry weight basis) ranged from 20.0-63.0% and 8.1-62.4%, respectively. There are

attempts to investigate the factors which affect the yield of algal biochar. Ronsse et al. (2013),

for instance, reported that the yield of algal biochar is lower than those of other

13
lignocellulosic biomass, namely pine wood, wheat straw, and green waste. The yield

decreased when the pyrolysis temperature and residence time were increased from 300 to

750 °C and from 10 to 60 min, respectively. Similar results were also observed by Tag et al.

(2016) who postulated that such observations could be due to the differences in the chemical

components between algal and terrestrial biomass, particularly for the composition of cell

wall. Besides that, the high ash content in algal samples might act catalytically during

pyrolysis to alter the distribution of products.

4.1.2 Higher heating value

The higher heating value (HHV) is the amount of heat released when the sample (fuel)

is combusted and the products have returned to a temperature of 25 °C. It is typically

measured using a bomb calorimeter in the laboratory. As shown in Table 4, biochars derived

from microalgae (7.6-23.0 MJ/kg) and macroalgae (5.2-21.2 MJ/kg) have comparable HHV.

In general, the HHV of algal biochars is lower than that of lignocellulosic biochars (which

has been reported to exceed 30 MJ/kg), is likely attributed to a lower carbon content but

higher ash content in algal biomass.

4.1.3 Surface properties

The physical properties of biochars, such as their surface area, porosity, and pore

volume, are routinely analyzed. The Brunauer-Emmet-Teller (BET) surface area of biochar

can be estimated using a surface area analyzer while a scanning electron microscope (SEM)

can be used to profile the surface topography and particle structure. The surface area of algal

biochar is generally low but some studies reported that an increase in the pyrolysis

temperature could result in a higher surface area (Bird et al., 2011). Ronsse et al. (2013)

found that the ash content in the biomass is negatively correlated with the surface area of the

resulting biochar. Wang et al. (2013) reported that the surface area of biochar derived from C.

vulgaris (2.4 m2/g) was deemed to be low relative to biochar obtained from lignocellulosic

14
biomass. Working with biochar derived from macroalgae, Roberts et al. (2015b) reported that

biochar produced from the macroalgae Eucheuma sp. has a significantly higher surface area

(30.03-34.82 m2/g) than those of other species which ranged from 1.29-8.87 m2/g.

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. (2013) observed that the particles of biochar derived from a green

microalga, Chlorella vulgaris, are compact and irregular, and these are different from the

structure of the feedstock before pyrolysis. Similar results were also reported for the biochar

derived from Chlamydomonas reinhardtii (Torri et al. (2011). These observations are in

contrast with biochars from lignocellulosic biomass, which normally retain their feedstock’s

structure. Table 4 summarizes the physical properties of biochars derived from several

microalgae and macroalgae species.

4.2 Chemical properties of algal biochar

4.2.1 Proximate and ultimate analysis

In proximate analysis, the moisture, volatile matter, ash, and fixed carbon content of

the samples are measured. Ultimate analysis (also termed as elemental analysis), on the other

hand, includes the determination of carbon, hydrogen, nitrogen, sulphur, and oxygen within

the samples. As shown in Table 5, algal biochars are typically low in C but high in N and

minerals (ash) contents. The low C content, in comparison to those derived from

lignocellulosic biomass, is characteristic of biochars produced from micro- and macroalgae.

4.2.2 Inorganic elemental analysis

Inorganic mineral content of algal biochar is typically determined using the

inductively coupled plasma atomic emission spectrometry (ICP-AES). Previous work has

provided insights into the abundance of various inorganic elements in algal biochar and

15
enables comparison with biochar derived from other feedstock. For example, Wang et al.

(2013) reported that biochar derived from C. vulgaris contained higher concentrations of

various trace elements, including P, K, Mg and Ca, compared to biochar from lignocellulosic

biomass, In a separate study, biochars produced from both red and brown seaweeds have high

concentrations of N (0.3-2.8%), P (0.5-6.6 g/kg), and K (5.1-119 g/kg) (Roberts et al., 2015b).

Results from both studies, when taken together, seem to suggest that algal biochar has good

prospects as fertilizers in view of their mineral contents.

4.2.3 pH

As shown in Table 5, all algal biochar was found to be alkaline. The pH of biochar

derived from macroalgae ranges from 7.6-13.7, whereas little information is available

regarding those derived from microalgae. The pH of biochar is found to be affected by the

pyrolysis temperature and the algal samples. According to the study by Ronsse et al. (2013),

the pH of biochar is correlated to the presence of oxygen functionalities in the biochar. Tag et

al. (2016) noted that the pH of algal biochar (8.7-13.7) increased with an increase in the

pyrolysis temperature (250-600 °C). A plausible explanation is the increase in the relative ash

content in the biochar caused a rise in the pH of the biochar especially under severe pyrolysis

conditions (Ronsse et al., 2013)

4.2.4 Cation exchange capacity

The cation exchange capacity (CEC) of biochar is a measure of the ability of biochar

to adsorb cation nutrients. In other words, a biochar with high CEC exerts beneficial effect

by preventing nutrient leaching in the soil. Roberts et al. (2015b) showed that seaweed

biochar samples have negligible or no exchangeable Al, but have high levels of the remaining

exchangeable cations such as Ca, K, Mg and Na. In another study by Tag et al. (2016), algal

biochar produced at different pyrolysis temperature was found to have higher CEC (25.6-52.6

cmol/kg) than that produced from vine pruning (32.2-61.0 cmol/kg) and orange pomace

16
(25.6-52.6 cmol/kg) under similar experimental conditions. As CEC correlated with the ash

content, it was suggested that the alkali and earth alkali metals in biomass promoted the

formation of O-containing surface functional groups in the resulting biochar (Cely et al.,

2015)

4.2.5 Functional groups

Fourier transform infrared spectroscopy (FTIR) spectral analysis is used to identify

the various functional groups in algal biochar. This technique can be used to compare the

chemical profiles of biochar produced under different conditions, as well as between the raw

biomass and the resulting biochar. The latter is exemplified by the work of Biswas et al.

(2017) whereby FT-IR spectra demonstrated that, following hydrothermal liquefaction, the

macromolecular crystal structure of the algae was disrupted and products (biochar and bio-oil)

were formed. Table 5 summarizes the chemical properties of biochar derived from some

microalgae and macroalgae species.

4.3 Factors affecting the properties of algal biochar

In general, the method and temperature of pyrolysis are two important factors that

determine the physical and chemical properties of biochar (Jindo et al. (2014); (Mukome et

al., 2013). The operating conditions such as heating rate, reaction vessel, chemical activation,

residence time, and highest treatment temperature (HTT) also influence the properties of

biochar produced. The HHT is regarded to have the greatest effect on the physical properties

of biochar produced (Mukome et al., 2013). In a recent study, Palanisamy et al. (2017)

reported that biochar of C. vulgaris prepared at higher temperatures (450-600 °C) contained a

higher proportion of organic matter (C, H and N) than those produced at lower temperatures.

In addition, the type of feedstock used may affect the chemical composition of the

resulting biochar, such as the ash content, pH, H/C ratio, surface area, as well as cation and

17
anion exchange capabilities (Sun et al., 2014; Tag et al., 2016). These factors are of great

importance when comes to the applications of biochar, particularly in the agriculture sector.

In a comparative analysis in the properties of biochar produced by algae, grass, manure,

wood, pomace, and nutshell, it was found that feedstock is a better predictor of variation in

the ash content and C/N ratio of biochar than pyrolysis temperature (Mukome et al. (2013).

However, when one feedstock is being considered, pyrolysis temperature is the best predictor

for the surface area.

In terms of biochar production from algal species, the proportion and composition of

nutrients of algal-derived biochar is likely influenced by a number of abiotic and biotic

factors like species, habitat (e.g. fresh, brackish or saline environment), and other factors. It

is evidenced by the findings of a number of researchers. Dealing with seaweed biochars,

Roberts et al. (2015b) found that biochars produced from red and brown seaweeds have

different elemental composition with biochars from red seaweeds have higher concentrations

of S and K and lower concentrations of C and H than biochars produced from the brown

seaweeds. There are also variations in most physicochemical characteristics of the biochar

between species collected from different locations especially the cation exchange capability

of the biochars. For instance, the exchangeable K in Undaria sp. (13-420 cmol/kg) and

Kappaphycus sp. (26-210 cmol/kg) differed by an order of magnitude depending on the

location. Bird et al. (2012) also noted a difference in the yield and ash content of the

Chadophora vagabunda with their earlier work (Bird et al., 2011). These may be attributed to

moving to pilot scale production with commercial pyrolysis equipment rather than controlled

laboratory production, purity of the materials collected in the field, and lesser control over

pyrolysis conditions.

4.4 Comparison of algal biochar with biochars derived from different feedstocks

18
It is generally accepted that the first critical step in determining the utility and

applications of algal biochar is the quantification of the properties of algal biochar and a

comparison with biochar produced from other commonly used terrestrial biomass feedstocks

(Bird et al., 2011). Studies thus far have pointed to the fact that biochars produced from algal

samples are fundamentally different from those produced from lignocellulosic feedstock

(Bird et al., 2012; Maddi et al., 2011). In general, biochars produced from various algal

species tend to have low carbon content, surface area and cation exchange capability but high

in pH, nitrogen, and extractable inorganic nutrients including P, K, Ca and Mg. In contrast,

lignocellulosic materials-derived biochars tend to have higher carbon contents and cation

exchange capabilities with pH values usually lower than 7, and significantly lower ash and

available nutrient contents (Jindo et al., 2014).

The difference in the chemical profiles between algal and other lignocellulosic

biochar gives important implications in their applications. The algal biochar is likely to

provide significant direct nutrient benefits to soils and crop productivity, and are likely to be

particularly useful for application on acidic soils. However, they are volumetrically less able

to provide the carbon sequestration benefits that can be provided with the high-carbon-

content lignocellulosic biochars. Therefore, Roberts et al. (2015b) proposed that a blending of

seaweed and lignocellulosic biochars could provide a soil ameliorant that combines a high

fixed C content with a mineral-rich substrate to enhance crop productivity.

5. Future challenges and opportunities

Algal biochar could influence the global carbon sequestration on mitigation of climate

change from the production of energy to the applications in environmental management. In

particular, the use of algal biochar as biosorbents in wastewater treatment should be studied

further. Nevertheless, it is important to take note of some of the challenges that may lie ahead,

19
such as cost effectiveness of the process and the fact that wastewater contains various

impurities that may interfere with the process. In addition, production of algal biochar via the

existing methods could be further enhanced with microwave-assisted technology, whereas the

additional energy consumption associated with the microwave-assisted process should be

well considered and minimized. Microwave-assisted technology in the production of algal

biochar is indeed a promising approach that deserves more attention in the future.

Furthermore, optimization of algal biochar productivity can also be achieved by employing a

continuous process.

In terms of the algal biomass to be used for biochar production, microalgae seem to be

a more promising feedstock when compared to macroalgae, in view of their advantages over

the latter, such as their rapid growth and the ease of cultivation and harvesting. Macroalgae,

on the other hand, require a larger area for cultivation and generally require a longer period of

cultivation time. The quality and quantity of macroalgal biomass are also greatly dependent

on varieties of abiotic and biotic factors. Moreover, microalgae are deemed to have higher

potential in biochar production due to their high nutrient contents which are suitable for the

use in agriculture. The recycling of microalgal residue after biofuel production for a

biorefinery concept is also feasible. Taken together, future research on different aspects

pertaining to microalgal biochar, such as production methods and newer applications, is

highly recommended in view of the scarcity of information on these aspects.

6. Conclusions

Algae biomass as a third generation feedstock can be utilized in biochar production in

the context of biorefinery. Pyrolysis, torrefaction or hydrothermal carbonization are suitable

processes to produce algal biochar of different composition and properties for varying further

applications. Microwave-assisted technology could be an interesting approach in enhancing

20
the current biochar production technologies. Characterization of algal biochar is important for

the understanding of their chemical and physical properties, which are useful for determining

their potential applications (such as fertilizer for agricultural purpose or adsorbent for water

treatment). Development of algal biochar technologies is expected to contribute to a more

sustainable environment in the future.

Acknowledgements

This study is supported by University of Malaya under the SATU Joint Research

Scheme (RU018J-2016, RU018L-2016, RU018O-2016 and RU018C-2016). The financial

support from Taiwan’s Ministry of Science and Technology (MOST) under grant numbers of

MOST 106-3113-E-006-011, 106-3113-E-006-004-CC2, 104-2221-E-006-227-MY3, and

103-2221-E-006-190-MY3 is also greatly appreciated.

References

1. Al-Hamamre, Z., Saidan, M., Hararah, M., Rawajfeh, K., Alkhasawneh, H.E., Al-
Shannag, M. 2017. Wastes and biomass materials as sustainable-renewable energy
resources for Jordan. Renewable and Sustainable Energy Reviews, 67, 295-314.
2. Alhashimi, H.A., Aktas, C.B. 2017. Life cycle environmental and economic
performance of biochar compared with activated carbon: A meta-analysis. Resources,
Conservation and Recycling, 118, 13-26.
3. Awad, Y.M., Lee, S.-E., Ahmed, M.B.M., Vu, N.T., Farooq, M., Kim, I.S., Kim, H.S.,
Vithanage, M., Usman, A.R.A., Al-Wabel, M., Meers, E., Kwon, E.E., Ok, Y.S. 2017.
Biochar, a potential hydroponic growth substrate, enhances the nutritional status and
growth of leafy vegetables. Journal of Cleaner Production, 156, 581-588.
4. Aysu, T., Abd Rahman, N.A., Sanna, A. 2016. Catalytic pyrolysis of Tetraselmis and
Isochrysis microalgae by nickel ceria based catalysts for hydrocarbon production.
Energy, 103, 205-214.
5. Aysu, T., Sanna, A. 2015. Nannochloropsis algae pyrolysis with ceria-based catalysts
for production of high-quality bio-oils. Bioresour Technol, 194, 108-16.
21
6. 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 Conversion and Management, 141, 163-170.
7. Bach, Q.-V., Chen, W.-H., Sheen, H.-K., Chang, J.-S. 2017b. Gasification kinetics of
raw and wet-torrefied microalgae Chlorella vulgaris ESP-31 in carbon dioxide.
Bioresource Technology.
8. Basu, P. 2010. Chapter 3 - Pyrolysis and Torrefaction. in: Biomass Gasification and
Pyrolysis, Academic Press. Boston, pp. 65-96.
9. Bird, M.I., Wurster, C.M., de Paula Silva, P.H., Bass, A.M., de Nys, R. 2011. Algal
biochar – production and properties. Bioresource Technology, 102(2), 1886-1891.
10. Bird, M.I., Wurster, C.M., De Paula Silva, P.H., Paul, N.A., De Nys, R. 2012. Algal
biochar: effects and applications. GCB Bioenergy, 4(1), 61-69.
11. 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 algae.
Bioresour Technol.
12. Boakye, P., Lee, C.W., Lee, W.M., Woo, S.H. 2016. The Cell Viability on Kelp and Fir
Biochar and the Effect on the Field Cultivation of Corn. Clean Technology, 22(1), 29-34.
13. 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
dimorphus. Renewable Energy.
14. Broch, A., Jena, U., Hoekman, S., Langford, J. 2014. Analysis of Solid and Aqueous
Phase Products from Hydrothermal Carbonization of Whole and Lipid-Extracted Algae.
Energies, 7(1), 62.
15. Brownsort, P.A. 2009. Biomass pyrolysis processes: review of scope, control and
variability. Edinburgh: UK Biochar Research Center.
16. 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
Research, 1(2), 185-193.
17. Cely, P., Gascó, G., Paz-Ferreiro, J., Méndez, A. 2015. Agronomic properties of
biochars from different manure wastes. Journal of Analytical and Applied Pyrolysis, 111,
173-182.
18. Chaiwong, K., Kiatsiriroat, T., Vorayos, N., Thararax, C. 2012. Biochar production from
freshwater algae by slow pyrolysis. Maejo International Journal of Science and
Technology, 6(2).

22
19. Chaiwong, K., Kiatsiriroat, T., Vorayos, N., Thararax, C. 2013. Study of bio-oil and bio-
char production from algae by slow pyrolysis. Biomass and Bioenergy, 56, 600-606.
20. Chang, Y.-M., Tsai, W.-T., Li, M.-H. 2015. Chemical characterization of char derived
from slow pyrolysis of microalgal residue. Journal of Analytical and Applied Pyrolysis,
111, 88-93.
21. Chen, W.-H. 2015. Microalgae oil: algae cultivation and harvest, algae residue
torrefaction and Diesel engine emissions tests. Aerosol and air quality research.
22. Chen, W.-H., Huang, M.-Y., Chang, J.-S., Chen, C.-Y. 2014a. Thermal decomposition
dynamics and severity of microalgae residues in torrefaction. Bioresource technology,
169, 258-264.
23. Chen, W.-H., Huang, M.-Y., Chang, J.-S., Chen, C.-Y. 2015a. Torrefaction operation
and optimization of microalga residue for energy densification and utilization. Applied
Energy, 154, 622-630.
24. Chen, W.-H., Huang, M.-Y., Chang, J.-S., Chen, C.-Y., Lee, W.-J. 2015b. An energy
analysis of torrefaction for upgrading microalga residue as a solid fuel. Bioresource
Technology, 185, 285-293.
25. Chen, W.-H., Lin, B.-J., Huang, M.-Y., Chang, J.-S. 2015c. Thermochemical conversion
of microalgal biomass into biofuels: a review. Bioresource technology, 184, 314-327.
26. Chen, W.-H., Peng, J., Bi, X.T. 2015d. A state-of-the-art review of biomass torrefaction,
densification and applications. Renewable and Sustainable Energy Reviews, 44, 847-866.
27. Chen, W.-H., Wu, Z.-Y., Chang, J.-S. 2014b. Isothermal and non-isothermal torrefaction
characteristics and kinetics of microalga Scenedesmus obliquus CNW-N. Bioresource
technology, 155, 245-251.
28. 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. Applied
Energy, 181, 110-119.
29. Chiodo, V., Zafarana, G., Maisano, S., Freni, S., Urbani, F. 2016. Pyrolysis of different
biomass: Direct comparison among Posidonia Oceanica, Lacustrine Alga and White-
Pine. Fuel, 164, 220-227.
30. Conti, R., Fabbri, D., Vassura, I., Ferroni, L. 2016. Comparison of chemical and
physical indices of thermal stability of biochars from different biomass by analytical
pyrolysis and thermogravimetry. Journal of Analytical and Applied Pyrolysis, 122, 160-
168.

23
31. Coronella, C.J., Yan, W., Reza, M.T., Vasquez, V.R. 2012. Method for wet torrefaction
of a biomass, Google Patents.
32. Deng, J., Wang, G.-j., Kuang, J.-h., Zhang, Y.-l., Luo, Y.-h. 2009. Pretreatment of
agricultural residues for co-gasification via torrefaction. Journal of Analytical and
Applied Pyrolysis, 86(2), 331-337.
33. Du, Z. 2013. Thermochemical conversion of microalgae for biofuel production,
University of Minnesota.
34. Ennis, C.J., Evans, A.G., Islam, M., Ralebitso-Senior, T.K., Senior, E. 2012. Biochar:
carbon sequestration, land remediation, and impacts on soil microbiology. Critical
reviews in environmental science and technology, 42(22), 2311-2364.
35. Erlach, B., Harder, B., Tsatsaronis, G. 2012. Combined hydrothermal carbonization and
gasification of biomass with carbon capture. Energy, 45(1), 329-338.
36. Foley, P.M., Beach, E.S., Zimmerman, J.B. 2011. Algae as a source of renewable
chemicals: opportunities and challenges. Green Chemistry, 13(6), 1399-1405.
37. Francavilla, M., Manara, P., Kamaterou, P., Monteleone, M., Zabaniotou, A. 2015.
Cascade approach of red macroalgae Gracilaria gracilis sustainable valorization by
extraction of phycobiliproteins and pyrolysis of residue. Bioresour Technol, 184, 305-13.
38. Gronnow, M.J., Budarin, V.L., Mašek, O., Crombie, K.N., Brownsort, P.A.,
Shuttleworth, P.S., Hurst, P.R., Clark, J.H. 2013. Torrefaction/biochar production by
microwave and conventional slow pyrolysis–comparison of energy properties. Gcb
Bioenergy, 5(2), 144-152.
39. Guo, W.-Q., Zheng, H.-S., Li, S., Du, J.-S., Feng, X.-C., Yin, R.-L., Wu, Q.-L., Ren, N.-
Q., Chang, J.-S. 2016. Removal of cephalosporin antibiotics 7-ACA from wastewater
during the cultivation of lipid-accumulating microalgae. Bioresource Technology, 221,
284-290.
40. Heilmann, S.M., Davis, H.T., Jader, L.R., Lefebvre, P.A., Sadowsky, M.J., Schendel,
F.J., Von Keitz, M.G., Valentas, K.J. 2010. Hydrothermal carbonization of microalgae.
Biomass and Bioenergy, 34(6), 875-882.
41. Hoekman, S.K., Broch, A., Robbins, C., Zielinska, B., Felix, L. 2013. Hydrothermal
carbonization (HTC) of selected woody and herbaceous biomass feedstocks. Biomass
Conversion and Biorefinery, 3(2), 113-126.
42. Inyang, M.I., Gao, B., Yao, Y., Xue, Y., Zimmerman, A., Mosa, A., Pullammanappallil,
P., Ok, Y.S., Cao, X. 2016. A review of biochar as a low-cost adsorbent for aqueous

24
heavy metal removal. Critical Reviews in Environmental Science and Technology, 46(4),
406-433.
43. Jindo, K., Mizumoto, H., Sawada, Y., Sanchez-Monedero, M.A., Sonoki, T. 2014.
Physical and chemical characterization of biochars derived from different agricultural
residues. Biogeosciences, 11(23), 6613-6621.
44. Johansson, C.L., Paul, N.A., de Nys, R., Roberts, D.A. 2016. Simultaneous biosorption
of selenium, arsenic and molybdenum with modified algal-based biochars. Journal of
Environmental Management, 165, 117-123.
45. Jung, K.-W., Jeong, T.-U., Kang, H.-J., Ahn, K.-H. 2016. Characteristics of biochar
derived from marine macroalgae and fabrication of granular biochar by entrapment in
calcium-alginate beads for phosphate removal from aqueous solution. Bioresource
Technology, 211, 108-116.
46. Kołtowski, M., Charmas, B., Skubiszewska-Zięba, J., Oleszczuk, P. 2017. Effect of
biochar activation by different methods on toxicity of soil contaminated by industrial
activity. Ecotoxicology and Environmental Safety, 136, 119-125.
47. Kubo, S. 2013. Nanostructured carbohydrate–derived carbonaceous materials. TANSO,
2013(258), 232-233.
48. Kumar, G., Shobana, S., Chen, W.-H., Bach, Q.-V., Kim, S.-H., Atabani, A.E., Chang,
J.-S. 2017. A review of thermochemical conversion of microalgal biomass for biofuels:
chemistry and processes. Green Chemistry, 19(1), 44-67.
49. Lei, H., Ren, S., Wang, L., Bu, Q., Julson, J., Holladay, J., Ruan, R. 2011. Microwave
pyrolysis of distillers dried grain with solubles (DDGS) for biofuel production.
Bioresource Technology, 102(10), 6208-6213.
50. Levine, R.B., Sierra, C.O.S., Hockstad, R., Obeid, W., Hatcher, P.G., Savage, P.E. 2013.
The use of hydrothermal carbonization to recycle nutrients in algal biofuel production.
Environmental Progress & Sustainable Energy, 32(4), 962-975.
51. Libra, J.A., Ro, K.S., Kammann, C., Funke, A., Berge, N.D., Neubauer, Y., Titirici, M.-
M., Fühner, C., Bens, O., Kern, J. 2011. Hydrothermal carbonization of biomass
residuals: a comparative review of the chemistry, processes and applications of wet and
dry pyrolysis. Biofuels, 2(1), 71-106.
52. Lu, Y., Levine, R.B., Savage, P.E. 2014. Fatty acids for nutraceuticals and biofuels from
hydrothermal carbonization of microalgae. Industrial & Engineering Chemistry
Research, 54(16), 4066-4071.

25
53. Maddi, B., Viamajala, S., Varanasi, S. 2011. Comparative study of pyrolysis of algal
biomass from natural lake blooms with lignocellulosic biomass. Bioresour Technol,
102(23), 11018-26.
54. Mehrabadi, A., Craggs, R., Farid, M.M. 2016. Pyrolysis of wastewater treatment high
rate algal pond (WWT HRAP) biomass. Algal Research.
55. Mohan, D., Sarswat, A., Ok, Y.S., Pittman, C.U. 2014. Organic and inorganic
contaminants removal from water with biochar, a renewable, low cost and sustainable
adsorbent–a critical review. Bioresource technology, 160, 191-202.
56. Molina, M., Zaelke, D., Sarma, K.M., Andersen, S.O., Ramanathan, V., Kaniaru, D.
2009. Reducing abrupt climate change risk using the Montreal Protocol and other
regulatory actions to complement cuts in CO2 emissions. Proceedings of the National
Academy of Sciences, 106(49), 20616-20621.
57. Mukome, F.N., Zhang, X., Silva, L.C., Six, J., Parikh, S.J. 2013. Use of chemical and
physical characteristics to investigate trends in biochar feedstocks. J Agric Food Chem,
61(9), 2196-204.
58. Neveux, N., Yuen, A.K., Jazrawi, C., Magnusson, M., Haynes, B.S., Masters, A.F.,
Montoya, A., Paul, N.A., Maschmeyer, T., de Nys, R. 2014. Biocrude yield and
productivity from the hydrothermal liquefaction of marine and freshwater green
macroalgae. Bioresour Technol, 155, 334-41.
59. Nhuchhen, D.R., Basu, P., Acharya, B. 2014. A comprehensive review on biomass
torrefaction. International Journal of Renewable Energy & Biofuels, 2014, 1-56.
60. Nizamuddin, S., Baloch, H.A., Griffin, G., Mubarak, N., Bhutto, A.W., Abro, R., Mazari,
S.A., Ali, B.S. 2017. An overview of effect of process parameters on hydrothermal
carbonization of biomass. Renewable and Sustainable Energy Reviews, 73, 1289-1299.
61. Norouzi, O., Jafarian, S., Safari, F., Tavasoli, A., Nejati, B. 2016. Promotion of
hydrogen-rich gas and phenolic-rich bio-oil production from green macroalgae
Cladophora glomerata via pyrolysis over its bio-char. Bioresource Technology, 219,
643-651.
62. Palanisamy, M., Mukund, S., Sivakumar, U., Karthikeyan, Sivasubramanian, V. 2017.
Bio-char production from micro algal biomass of Chlorella vulgaris. PHYKOS, 47(1),
99-104.
63. Park, S.H., Cho, H.J., Ryu, C., Park, Y.-K. 2016. Removal of copper(II) in aqueous
solution using pyrolytic biochars derived from red macroalga Porphyra tenera. Journal
of Industrial and Engineering Chemistry, 36, 314-319.

26
64. Rashid, N., Rehman, M.S.U., Han, J.-I. 2013. Recycling and reuse of spent microalgal
biomass for sustainable biofuels. Biochemical engineering journal, 75, 101-107.
65. Roberts, D.A., Cole, A.J., Paul, N.A., de Nys, R. 2015a. Algal biochar enhances the re-
vegetation of stockpiled mine soils with native grass. Journal of Environmental
Management, 161, 173-180.
66. Roberts, D.A., Paul, N.A., Dworjanyn, S.A., Bird, M.I., de Nys, R. 2015b. Biochar from
commercially cultivated seaweed for soil amelioration. Sci Rep, 5, 9665.
67. Ronsse, F., van Hecke, S., Dickinson, D., Prins, W. 2013. Production and
characterization of slow pyrolysis biochar: influence of feedstock type and pyrolysis
conditions. GCB Bioenergy, 5(2), 104-115.
68. Roy, P., Dias, G. 2017. Prospects for pyrolysis technologies in the bioenergy sector: A
review. Renewable and Sustainable Energy Reviews, 77, 59-69.
69. Sarkar, O., Agarwal, M., Naresh Kumar, A., Venkata Mohan, S. 2015. Retrofitting
hetrotrophically cultivated algae biomass as pyrolytic feedstock for biogas, bio-char and
bio-oil production encompassing biorefinery. Bioresour Technol, 178, 132-8.
70. Shukla, S., Gita, S., Bharti, V., Bhuvaneswari, G., Wikramasinghe, W. 2017.
Atmospheric Carbon Sequestration Through Microalgae: Status, Prospects, and
Challenges. in: Agro-Environmental Sustainability, Springer, pp. 219-235.
71. Shuttleworth, P., Budarin, V., Gronnow, M., Clark, J.H., Luque, R. 2012. Low
temperature microwave-assisted vs conventional pyrolysis of various biomass
feedstocks. Journal of Natural Gas Chemistry, 21(3), 270-274.
72. Smith, A.M., Ross, A.B. 2016. Production of bio-coal, bio-methane and fertilizer from
seaweed via hydrothermal carbonisation. Algal Research, 16, 1-11.
73. Sohi, S., Loez-Capel, E., Krull, E., Bol, R. 2009. Biochar’s roles in soil and climate
change: A review of research needs. CSIRO Land and Water Science Report, 5(09), 1-57.
74. Spokas, K.A., Novak, J.M., Stewart, C.E., Cantrell, K.B., Uchimiya, M., DuSaire, M.G.,
Ro, K.S. 2011. Qualitative analysis of volatile organic compounds on biochar.
Chemosphere, 85(5), 869-882.
75. Suali, E., Sarbatly, R. 2012. Conversion of microalgae to biofuel. Renewable and
Sustainable Energy Reviews, 16(6), 4316-4342.
76. Suganya, T., Varman, M., Masjuki, H., Renganathan, S. 2016. Macroalgae and
microalgae as a potential source for commercial applications along with biofuels
production: A biorefinery approach. Renewable and Sustainable Energy Reviews, 55,
909-941.

27
77. Sun, H., Lu, H., Chu, L., Shao, H., Shi, W. 2017. Biochar applied with appropriate rates
can reduce N leaching, keep N retention and not increase NH3 volatilization in a coastal
saline soil. Science of The Total Environment, 575, 820-825.
78. Sun, Y., Gao, B., Yao, Y., Fang, J., Zhang, M., Zhou, Y., Chen, H., Yang, L. 2014.
Effects of feedstock type, production method, and pyrolysis temperature on biochar and
hydrochar properties. Chemical Engineering Journal, 240, 574-578.
79. Tag, A.T., Duman, G., Ucar, S., Yanik, J. 2016. Effects of feedstock type and pyrolysis
temperature on potential applications of biochar. Journal of Analytical and Applied
Pyrolysis, 120, 200-206.
80. Tekin, K., Karagöz, S., Bektaş, S. 2014. A review of hydrothermal biomass processing.
Renewable and sustainable Energy reviews, 40, 673-687.
81. Titirici, M.-M., White, R.J., Falco, C., Sevilla, M. 2012. Black perspectives for a green
future: hydrothermal carbons for environment protection and energy storage. Energy &
Environmental Science, 5(5), 6796-6822.
82. Torri, C., Samorì, C., Adamiano, A., Fabbri, D., Faraloni, C., Torzillo, G. 2011.
Preliminary investigation on the production of fuels and bio-char from Chlamydomonas
reinhardtii biomass residue after bio-hydrogen production. Bioresource Technology,
102(18), 8707-8713.
83. Tripathi, M., Sahu, J.N., Ganesan, P. 2016. Effect of process parameters on production
of biochar from biomass waste through pyrolysis: A review. Renewable and Sustainable
Energy Reviews, 55, 467-481.
84. Uemura, Y., Matsumoto, R., Saadon, S., Matsumura, Y. 2015. A study on torrefaction of
Laminaria japonica. Fuel Processing Technology, 138, 133-138.
85. Vassilev, S.V., Vassileva, C.G. 2016. Composition, properties and challenges of algae
biomass for biofuel application: An overview. Fuel, 181, 1-33.
86. Wan, Y., Chen, P., Zhang, B., Yang, C., Liu, Y., Lin, X., Ruan, R. 2009. Microwave-
assisted pyrolysis of biomass: Catalysts to improve product selectivity. Journal of
Analytical and Applied Pyrolysis, 86(1), 161-167.
87. Wang, K., Brown, R.C., Homsy, S., Martinez, L., Sidhu, S.S. 2013. Fast pyrolysis of
microalgae remnants in a fluidized bed reactor for bio-oil and biochar production.
Bioresource Technology, 127, 494-499.
88. Wang, N., Tahmasebi, A., Yu, J., Xu, J., Huang, F., Mamaeva, A. 2015. A Comparative
study of microwave-induced pyrolysis of lignocellulosic and algal biomass. Bioresource
Technology, 190, 89-96.

28
89. Wiedner, K., Rumpel, C., Steiner, C., Pozzi, A., Maas, R., Glaser, B. 2013. Chemical
evaluation of chars produced by thermochemical conversion (gasification, pyrolysis and
hydrothermal carbonization) of agro-industrial biomass on a commercial scale. Biomass
and Bioenergy, 59, 264-278.
90. Wu, K.-T., Tsai, C.-J., Chen, C.-S., Chen, H.-W. 2012. The characteristics of torrefied
microalgae. Applied energy, 100, 52-57.
91. Xiao, L.-P., Shi, Z.-J., Xu, F., Sun, R.-C. 2012. Hydrothermal carbonization of
lignocellulosic biomass. Bioresource Technology, 118, 619-623.
92. Xu, Q., Qian, Q., Quek, A., Ai, N., Zeng, G., Wang, J. 2013. Hydrothermal
carbonization of macroalgae and the effects of experimental parameters on the
properties of hydrochars. ACS Sustainable Chemistry & Engineering, 1(9), 1092-1101.
93. Yan, W., Acharjee, T.C., Coronella, C.J., Vásquez, V.R. 2009. Thermal pretreatment of
lignocellulosic biomass. Environmental Progress & Sustainable Energy, 28(3), 435-440.
94. Yanik, J., Stahl, R., Troeger, N., Sinag, A. 2013. Pyrolysis of algal biomass. Journal of
Analytical and Applied Pyrolysis, 103, 134-141.
95. Yao, C., Wu, P., Pan, Y., Lu, H., Chi, L., Meng, Y., Cao, X., Xue, S., Yang, X. 2016.
Evaluation of the integrated hydrothermal carbonization-algal cultivation process for
enhanced nitrogen utilization in Arthrospira platensis production. Bioresource
Technology, 216, 381-390.
96. Yin, C. 2012. Microwave-assisted pyrolysis of biomass for liquid biofuels production.
Bioresource technology, 120, 273-284.
97. Yuan, T., Tahmasebi, A., Yu, J. 2015. Comparative study on pyrolysis of lignocellulosic
and algal biomass using a thermogravimetric and a fixed-bed reactor. Bioresource
Technology, 175, 333-341.
98. Yuan, Y., Macquarrie, D.J. 2015. Microwave assisted step-by-step process for the
production of fucoidan, alginate sodium, sugars and biochar from Ascophyllum
nodosum through a biorefinery concept. Bioresour Technol, 198, 819-27.
99. Zeraatkar, A.K., Ahmadzadeh, H., Talebi, A.F., Moheimani, N.R., McHenry, M.P. 2016.
Potential use of algae for heavy metal bioremediation, a critical review. Journal of
environmental management, 181, 817-831.
100. Zheng, H., Guo, W., Li, S., Chen, Y., Wu, Q., Feng, X., Yin, R., Ho, S.-H., Ren, N.,
Chang, J.-S. 2017. Adsorption of p-nitrophenols (PNP) on microalgal biochar: Analysis
of high adsorption capacity and mechanism. Bioresource Technology.

29
Table 1. The pyrolysis processes used for algal biochar production

Types of Biomass Temperature


Biochar production References
process feedstock (℃)
Slow Chlorella-based 300-700  56.3% at 300 °C (Chang et al.,
pyrolysis algal residue  66.2% at 500 °C 2015)
 65.0% at 700 °C
 High concentration of
nitrogen and other
inorganic elements

Slow Brown Laminaria 200-800  78.34% at 200 ℃ (Jung et al.,


pyrolysis japonica  63.64% at 400 ℃ 2016)
macroalgae  37.96% at 600 ℃
 27.05% at 800 ℃

Fixed-bed Chlorella 300-900  19.3-43.46% of biochar (Yuan et al.,


pyrolysis vulgaris yield on different 2015)
temperatures

Fixed-bed Chlamydomonas 350  Nitrogen-rich biochar (Torri et al.,


pyrolysis reinhardtii  Largest fraction in term 2011)
of mass, 44 ± 1% w/w
mass yield of biochar

Fixed-bed Scenedesmus 300-600  Surface area of biochar (Bordoloi et


pyrolysis dimorphus increased from 1.72 to al., 2016)
123 m2/g when
temperature increased
from 300-500 °C;
reduced to 89 m2/g at
600 °C
 The recalcitrance of
biochar increased from
0.62 to 0.76 with
increasing temperature

Fixed-bed Seaweed 250-600  Biochar energy yield of (Tag et al.,


pyrolysis 61.50% at 600 ℃ to 2016)
93.95% at 250 ℃

Fixed-bed Green macroalgae 400-600  44 wt% yield at 400 ℃ (Norouzi et


pyrolysis Cladophora  40 wt% yield at 500 ℃ al., 2016)
glomerata  39 wt% yield at 600 ℃

31
Fluidized- Defatted 500  31% of biochar yield (Wang et al.,
bed fast Chlorella  Energy recovery of 2013)
pyrolysis vulgaris algal biomass in bio-oil
and biochar is 94%
 High inorganic biochar
content

Fluidized- Laminaria 500  29-36% of biochar yield (Yanik et al.,


bed digitata, Fucus  Production of biochar 2013)
pyrolysis serratus and mix with lower heating
macroalgae value due to higher ash
species from content of 41-52%
Black sea

Stepwise Wastewater 300-500  High amount of biochar (Mehrabadi et


and non- treatment high with more than 50 wt% al., 2016)
stepwise rate algal pond of the initial biomass
pyrolysis (WWT HRAP) was produced under
biomass both heating regimes

Microwave Macroalgae 240-400  54.8% of biochar yield (Shuttleworth


pyrolysis et al., 2012)

32
Table 2. Comparison of the characteristics of dry and wet torrefaction (Yan et al., 2009)

Torrefaction
Characteristics
Dry Wet (Hydrothermal)
Temperature 200-300 ℃ 180-260 ℃
Media Inert nitrogen gas Hot compressed water
Pressure Atmospheric pressure 200-700 psi
Residence time 80 min 5 min
Cooling process Flowing nitrogen; indirect water Immerse into ice bath rapidly
cooling
Additional processes - Filtration and evaporation
Energy density Lower Higher

33
Table 3. Algal hydrochar production using hydrothermal carbonization (HTC)

Temperature Solid mass yield


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

Macroalgae
(Xu et al.,
Sargassum horneri 180-210 2-16 hours 32.7-52.3
2013)
(Smith &
Laminaria digitata 200-250 1 hour 18.4-21.8
Ross, 2016)
(Smith &
Laminaria hyperborean 200-250 1 hour 23.6-39.0
Ross, 2016)
(Smith &
Alaria esculenta 200-250 1 hour 23.7-30.0
Ross, 2016)
Table 4. Physical properties of biochars derived from microalgal and macroalgal biomass

Pyrolysis Biochar BET


HHV
Feedstock temperature and yield surface area Reference
(MJ/kg)
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 350 °C, 20 min 44.0 13.0 n.d. Torri et al.
reinhardtii (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 500 °C, 20 min 34.0 n.d. n.d. Conti et al.
communis (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.

34
(2011)
Nannochloropsis sp. 400, 500, and 24.8- 14.8 n.d. Aysu and
600 °C, 60 min 33.5 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 41.3- n.d. n.d. Chiodo et al.
600 °C, 45 min 48.3 (2016)
Mixed consortia 500, 600, and 33.0- n.d. n.d. Sarkar et al.
800 °C, 1, 2, and 63.0 (2015)
3h

Macroalgae

Ascophyllum nodosum Microwave 21.4 21.2 n.d. Yuan and


assisted thermal Macquarrie
treatment (2015)
Cladophora sp. 550 °C, 60 min 31.0 16.7 n.d. Chaiwong et
al. (2012)
Cladophora 250-400 °C 67.0 n.d. 8.3 Bird et al.
vagabunda (2012)
Cladophora coelothrix Hydrothermal 10.4 18.3 n.d. Neveux et al.
liquefaction (2014)
Cladophora Hydrothermal 18.7 15.1 n.d. Neveux et al.
vagabunda liquefaction (2014)
Chaetomorpha linum Hydrothermal 8.4 20.5 n.d. Neveux et al.
liquefaction (2014)
Derbesia tenuissima Hydrothermal 8.1 9.1 n.d. Neveux et al.
liquefaction (2014)
Eucheuma sp. 450 °C, 60 min 57.2- 14.6- 30.0-34.8 Roberts et al.
61.7 17.2 (2015b)
Gracilaria sp. 450 °C, 60 min 59.8- 11.1- 2.0-3.6 Roberts et al.
61.8 16.1 (2015b)
Gracilaria gracilis 400, 500, and n.d. 11.4 n.d. Francavilla et
600 °C al. (2015)
Gracilaria gracilis 400, 500, and 26-32 14.5 n.d. Francavilla et
residue 600 °C al. (2015)
Kappaphycus sp. 450 °C, 60 min 54.1- 13.0- 2.2-2.8 Roberts et al.
59.2 17.8 (2015b)
Oedogonium sp. Hydrothermal 10.2 5.2 n.d. Neveux et al.
liquefaction (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- 11.4- 1.3-8.5 Roberts et al.
49.7 14.8 (2015b)
Sargassum sp. 450 °C, 60 min 49.0- 11.8- 2.5-7.5 Roberts et al.
61.9 13.5 (2015b)
Ulva flexuousa 300-500 °C, 60 n.d. n.d. 1.2-4.3 Bird et al.
min (2011)

35
Ulva ohnoi Hydrothermal 12.1 4.0 n.d. Neveux et al.
liquefaction (2014)
Undaria sp. 450 °C, 60 min 60.3- 10.7- 1.3-8.9 Roberts et al.
62.4 14.7 (2015b)
Mixture of 250-400 °C 74.5 n.d. n.d. Bird et al.
Cladophora (2012)
coelothrix,
Chaetomorpha indica,
Ulva flexuosa
Kelp n.a. n.d. n.d. 26.6 Boakye et al.
(2016)
Spray-dried algae 300, 450, 600, 19.3- 8.2-9.2 14-19 Ronsse et al.
and 750 °C, 10 72.8 (2013)
and 60 min
n.a., not available; n.d., not determined.

36
Table 5. Chemical properties of biochars derived from microalgal and macroalgal biomass

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


Biochar Volatile Fixed pH Reference
Moisture Ash C H N S O
matter carbon

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 ‒ 22.0 36.8 41.0 51.0 3.0 7.0 0.1 14.0 ‒ Conti et al. (2016)
communis
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.

37
(2012)
Tetraselmis sp. ‒ ‒ ‒ ‒ 48.1 1.7 4.3 ‒ 46.0 ‒ Aysu et al. (2016)
Lacustrine algae ‒ ‒ ‒ ‒ 27.0- 0.8- 0.7- ‒ 67.1- 10.05 Chiodo et al.
30.5 1.1 1.3 71.5 (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 ‒ ‒ 47.0 ‒ 20.3 1.2 1.7 ‒ ‒ 9.1 Bird et al. (2011)
patentiramea
Cladophora ‒ ‒ 54.2 ‒ 21.8 1.2 2.0 ‒ ‒ 9.9 Bird et al. (2011)
vagabunda
Cladophora ‒ ‒ ‒ ‒ 36.1 3.9 2.6 0.3 20.6 ‒ Neveux et al.
vagabunda (2014)
Cladophora ‒ ‒ 47.0 ‒ 20.3 1.2 1.7 ‒ ‒ 9.1 Bird et al. (2011)
patentiramea
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)

38
Eucheuma sp. ‒ ‒ ‒ ‒ 23.7- 1.2- 0.7- 7.0- 20.6- 8.2-8.6 Roberts et al.
25.6 1.8 0.8 9.3 24.9 (2015b)
Gracilaria sp. ‒ ‒ ‒ ‒ 24.5- 1.5- 1.3- 2.7- 16.5- 7.6-8.1 Roberts et al.
30.9 2.2 2.8 4.4 19.8 (2015b)
Gracilaria gracilis 4.1 ‒ 47.4 ‒ 36.8 0.2 2.3 2.6 10.6 ‒ Francavilla et al.
(2015)
Gracilaria gracilis 5.4 ‒ 30.7 ‒ 41.2 2.2 2.3 1.3 22.3 ‒ Francavilla et al.
residue (2015)
Kappaphycus sp. ‒ ‒ ‒ ‒ 22.2- 1.1- 0.3- 5.5- 15.6- 8.8-9.0 Roberts et al.
31.3 2.1 0.7 6.8 23.8 (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- 1.9- 2.2- 1.0- 16.4- 11.0- Roberts et al.
35.0 2.4 2.4 1.6 18.4 11.2 (2015b)
Sargassum sp. ‒ ‒ ‒ ‒ 28.9- 2.0- 1.0- 0.9- 15.3- 10.1- Roberts et al.
29.1 2.1 1.1 2.8 18.2 10.8 (2015b)
Ulva flexuousa ‒ ‒ 26.5- ‒ 22.3- 1.1- 4.0- ‒ ‒ 8.0- Bird et al. (2011)
42.6 30.9 3.2 7.7 10.1
Ulva ohnoi ‒ ‒ ‒ ‒ 9.6 2.3 0.9 16.3 35.1 ‒ Neveux et al.
(2014)
Undaria sp. ‒ ‒ ‒ ‒ 27.3- 1.7- 2.3- 0.6- 14.1- 9.9- Roberts et al.
34.8 2.8 2.4 0.8 15.6 10.9 (2015b)
Mixture ‒ 11.6-54.2 22.9- 22.9-45.7 36.6- 1.4- 1.4- 0-0.8 9.3- 8.7- Tag et al. (2016)
42.7 45.3 3.7 2.1 34.7 13.7
Kelp ‒ ‒ ‒ ‒ 27.4 0.6 1.28 ‒ 12.1 ‒ Boakye et al.
(2016)
Spray-dried algae ‒ 3.9-70.0 46.3- 30.0-96.1 62.7- 1.4- ‒ ‒ ‒ 4.9- Ronsse et al.
76.4 90.6 7.2 12.5 (2013)
‒: not determined

39
Figure 1. Algal biomass production for renewable energy production and carbon
sequestration

40
Figure 2. General torrefaction process for algal biochar production

41
Highlights

 Conventional and newly developed methods for algal biochar synthesis are discussed

 Characterization of algal biochars is presented to justify their applications

 Future directions on the development of algal biochar technologies are proposed

42
Graphical abstract

43

You might also like