You are on page 1of 42

Accepted Manuscript

Review

Pyrolysis characteristics and kinetics of microalgae via thermogravimetric anal-


ysis (TGA): A state-of-the art review

Quang-Vu Bach, Wei-Hsin Chen

PII: S0960-8524(17)30991-4
DOI: http://dx.doi.org/10.1016/j.biortech.2017.06.087
Reference: BITE 18324

To appear in: Bioresource Technology

Received Date: 24 May 2017


Revised Date: 15 June 2017
Accepted Date: 16 June 2017

Please cite this article as: Bach, Q-V., Chen, W-H., Pyrolysis characteristics and kinetics of microalgae via
thermogravimetric analysis (TGA): A state-of-the art review, Bioresource Technology (2017), doi: http://dx.doi.org/
10.1016/j.biortech.2017.06.087

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.
Pyrolysis characteristics and kinetics of microalgae via thermogravimetric analysis

(TGA): A state-of-the art review

Quang-Vu Bach a, Wei-Hsin Chen b,*


a
Sustainable Management of Natural Resources and Environment Research Group, Faculty of
Environment and Labour Safety, Ton Duc Thang University, Ho Chi Minh City, Vietnam
b
Department of Aeronautics and Astronautics, National Cheng Kung University, Tainan 701, Taiwan
* Corresponding author
Tel: +886-6-200 4456; Fax: +886-6-238 9940
E-mail: weihsinchen@gmail.com; chenwh@mail.ncku.edu.tw (Wei-Hsin Chen)
bachquangvu@tdt.edu.vn (Quang-Vu Bach)

Abstract

Pyrolysis is a promising route for biofuels production from microalgae at moderate

temperatures (400 to 600 °C) in an inert atmosphere. Depending on the operating conditions,

pyrolysis can produce biochar and/or bio-oil. In practice, knowledge for thermal decomposition

characteristics and kinetics of microalgae during pyrolysis is essential for pyrolyzer design and

pyrolysis optimization. Recently, the pyrolysis kinetics of microalgae has become a crucial

topic and received increasing interest from researchers. Thermogravimetric analysis (TGA) has

been employed as a proven technique for studying microalgae pyrolysis in a kinetic control

regime. In addition, a number of kinetic models have been applied to process the TGA data for

kinetic evaluation and parameters estimation. This paper aims to provide a state-of-the art

review on recent research activities in pyrolysis characteristics and kinetics of various

microalgae. Common kinetic models predicting the thermal degradation of microalgae are

examined and their pros and cons are illustrated.

Keywords: Microalgal biomass; Pyrolysis and torrefaction; Kinetics; Thermogravimetric

analysis; Biochar and bio-oil.

1
Nomenclature

Abbreviations

TGA Thermogravimetric analysis


DTG Differential thermogravimetric
FWO Flynn-Wall-Ozawa model
KAS Kissinger-Akahira-Sunose model
DAEM Distributed activation energy model
Symbols

Pre-exponential factor (s-1)


Contribution factor
Activation energy (kJ/mol)
Mean activation energy in DAEM (kJ/mol)
or Reaction rate constant (s-1)
Sample mass at any time (g)
Initial sample mass (g)
Final residual mass (g)
Universal gas constant (=8.314 J.mol-1.K-1)
Absolute temperature (K)
Conversion time (s)
Mass of released volatiles at any time (g)
Total mass of released volatiles (g)
Conversion degree
Heating rate (K/min)
Deviation of activation energy in DAEM (kJ/mol)

2
1. Introduction

Since the industrial revolution occurred in eighteenth century, a considerable amount of

coal has been burned for heat and power generation as well as industrial applications (Du &

Chen, 2006). The consumption of the carbon-based fuels leads to the mass emissions of

anthropogenic carbon dioxide into the atmosphere. Nowadays, rapid accumulation of carbon

dioxide in the atmosphere and the accompanied problems such as global warming and climate

change have become serious obstacles for environment sustainability (Janković & Schultz,

2017). For this reason, how to efficiently develop renewable energy and sustainable alternatives

to fossil fuels for mitigating the global warming is receiving a great deal of attention. A number

of renewable sources such as solar, wind, biomass, geothermal, marine, and hydroelectric

energies can be employed for green energy (Chen et al., 2012b). Solar energy and wind energy

have a substantial growth lately; however, electricity produced from solar and wind energy is

subject to climate and location, and is relatively difficult to be stored. In contrast, biomass

grows worldwide, which is not restricted geometrically and by the weather, and can be easily

stored and delivered after it is harvested. In addition, carbon stored in biomass completely

comes from carbon dioxide in the atmosphere. Consequently, biofuels derived from biomass

are considered as carbon-neutral fuels. If biochar produced from biomass is used for soil

amendment and remediation, it is even termed carbon negative material (Glaser et al., 2009) in

that biochar is mostly not digestible to microorganisms and a biochar-based soil amendment

could serve as a permanent carbon-sequestration agent in soils/subsoil earth layers for

thousands of years (Lee et al., 2010).

Biomass is currently the fourth largest primary energy source in the world, behind coal,

petroleum and natural gas (Chen et al., 2012b). According to the types of biomass feedstock,

biofuels can be categorized into first, second, third, and fourth generation biofuels. Biofuels

3
produced from food crops, say, sucrose- and starch-derived bioethanol, belong to first

generation biofuels (Chen et al., 2015a). Second generation biofuels are produced from

non-food crops such as lignocellulosic biomass. Biofuels using algal biomass such as

microalgae and macroalgae as feedstocks are termed third generation biofuels. As for fourth

generation biofuels, the approach deals with the metabolic engineering of algae from oxygenic

photosynthetic microorganisms (Lü et al., 2011). Obviously, third and fourth generation

biofuels involve algae-to-biofuel technology so that algal biomass is a promising substitute for

biofuel production.

Algal biomass possesses a number of advantages over lignocelluloses, including abundant

distribution, fast growth, high production rate, and high photosynthetic or carbon fixing

efficiency. Moreover, algal biomass is cultivated in water, even using wastewater as nutrient

sources (Chiu et al., 2015), and thus requires no arable land (Bach et al., 2017a). To obtain

biofuels from algal biomass through thermochemical conversion, a variety of routes can be

utilized, depending on the requirements of product phases. For example, biochar can be

obtained from torrefaction and pyrolysis, and bio-oil can be produced via pyrolysis and

liquefaction (Chen et al., 2015a). Regarding biochar production, torrefaction biochar is mainly

applied for solid fuels, whereas pyrolysis biochar is utilized for other purposes such as

activated carbon, soil enhancer, fertilizer, etc. (Manyà, 2012).

As far as microalgae pyrolysis is concerned, the feedstock is thermally decomposed in a

moderate-temperature and oxygen-free environment. The pressure in the reactor is normally

one atmosphere and the temperature is usually between 400 and 600 °C (Chen et al., 2015a).

Torrefaction is also operated in an oxygen-free environment, but at lower temperatures, namely,

200-300 °C (Chen et al., 2015b). Consequently, torrefaction is termed mild pyrolysis, and

broadly speaking, pertains to pyrolysis technology. Compared to combustion, gasification, and

4
liquefaction, it appears that pyrolysis is a more flexible process for producing biofuels in that

main products can be solid biochar or liquid bio-oil, depending on process operating conditions

such as heating rate, reaction temperature, residence time, and feedstock size.

To understand the fundamental pyrolysis characteristics of microalgae, thermogravimetric

analysis (TGA) has been widely performed. It is known that the reaction rate is highly related to

the reactor design and practical pyrolysis operation. Therefore, the chemical kinetics is a crucial

topic in microalgae pyrolysis and has been extensively studied. Over the past several decades, a

number of kinetic theories from TGA have been conducted to predict the thermal

decomposition rate of microalgal biomass during pyrolysis and to aid pyrolyzer design.

Nevertheless, the examination of published literature suggests that the review on the chemical

kinetics of microalgae pyrolysis through TGA remains insufficient so far. For this reason, this

paper aims to provide a comprehensive overview on recent research and development in

pyrolysis characteristics and kinetics using microalgae as feedstocks. A variety of models,

including single reaction, multiple paraellel reaction, series or consecutive rection, and

distributed activation energy models, in predicting the thermal degradation behaviors of

components in microalgae will be illustrated, and their pros and cons will be discussed. This

review is conducive to biochar and bio-oil production, biomass thermal degradation prediction,

and experiment and reactor designs.

2. Composition and thermal characteristic of microalgae

2.1. Microalgae and their chemical composition

Microalgae are photosynthetic unicellular or simple-multicellular microorganisms, and

are one of the earth’s most important natural resources (Sambusiti et al., 2015). They account

for approximately 50% of global photosynthetic activity (Chiu et al., 2015). Because some

5
microalgae double in number within hours, they have a short harvesting cycle, say, less than 10

days (Razeghifard, 2013). In recent years, the cultivation of microalgae has received a great

deal of attention for pharmaceutical, nutraceutical, cosmetic, aquiculture, and biorefinery

purposes (Zhu, 2015). In addition to growth in natural environments, microalgae can also be

artificially and commercially cultivated in freshwater, marine, and wastewater systems within

open ponds (raceways) and closed photobioreactors (Sambusiti et al., 2015). Certain

microalgae can tolerate and adapt to a wide variety of environmental conditions, and can be

produced all year round. While microalgae are cultivated, CO2 in flue gas can be used as a

carbon source for their growth. Moreover, microalgae are able to mitigate CO2 from 10 to 50

times higher than terrestrial plants (Raheem et al., 2015). As a consequence, microalgae-based

CO2 biological fixation is a promising means to mitigate anthropogenic CO2 emissions.

Microalgae can be grouped into prokaryotic microalgae (cyanobacteria Chloroxybacteria),

eukaryotic microalgae (green algae Chlorophyta), red algae (Rhodophyta), and diatoms

(Bacillariophta) (Sambusiti et al., 2015). Unlike lignocellulosic biomass which is mainly

composed of cellulose (40–60 wt%), hemicellulose (20–40 wt%), and lignin (10–25 wt%)

(Yang et al., 2007), the main chemical composition in microalgae consists of carbohydrates

(8–30 wt%), proteins (40–60 wt%), and lipids (5–60 wt%) (Uggetti et al., 2014); other valuable

components such as pigments, anti-oxidants, fatty acids, and vitamins are also contained.

Chemical composition of microalgae is high variable, and depends mainly on species,

environmental conditions, and cultivation methods (Sambusiti et al., 2015). Carbohydrates, the

fermentable sugars, in microalgae are potential feedstocks for bioethanol production

(Sirajunnisa & Surendhiran, 2016). Some microalgal cell walls are composed of cellulose,

mannans, xylans, and sulfated glycans. These polysaccharides can be chemically or

enzymatically broken down into simple sugars and then converted into bioethanol. Certain

6
oleaginous microalgae pertain to lipid-rich species in nature and can be cultivated at optimal

conditions to accumulate substantial quantities of lipids, making high oil yields

(5,000–100,000 L ha−1 year−1) (Singh & Olsen, 2011). These lipids in microalgae can be

extracted and converted into fatty acid methyl ester (FAME) as biodiesel through

transesterification (Mata et al., 2010). After microalgae undergoing oil-extraction, proteins in

solid residues become valuable co-products and can be used as feeds or fertilizers (Brennan &

Owende, 2010). The elemental and chemical compositions of some common microalgae are

tabulated in Table 1.

2.2. Thermal degradation characteristics of microalgae

The recognition of the thermal degradation of carbohydrates, proteins, and lipids plays a

crucial role in investigating the pyrolytic characteristics of microalgae. To figure out the

thermal decomposition behavior of microalgae, the thermogravimetric analysis (TGA) and

derivative thermogravimetric (DTG) curves of three different microalgal species, consisting of

Scenedesmus obliquus (S. obliquus) CNW-N (Chen et al., 2014), Chlorella vulgaris (C.

vulgaris) ESP-31 (Bach & Chen, 2017), and Chlamydomonas sp. (C. sp.) JSC4 (Nakanishi et

al., 2014), are shown in Figure 1 where the temperature is in the range of 105-800 °C (ignoring

the drying process), the heating rate is 20 °C min-1, and nitrogen at a flow rate of 100 mL min-1

(STP) is used to sweep the samples.

It can be seen that the microalga S. obliquus CNW-N has a smaller weight loss within the

investigated temperature range when compared with C. sp. JSC4 and C. vulgaris ESP-31

(Figure 1a), revealing that the former is relatively thermally resistant in nature. The DTG

curves of S. obliquus CNW-N, C. vulgaris ESP-31, and C. sp. JSC4 are characterized by a

single, two, and three peaks, respectively (Figure 1b). Their biggest peaks, corresponding to the

main pyrolysis or devolatilization process, are located at around 320, 290, and 280 °C,

7
respectively. These peaks triggered are due to the thermal decomposition of carbohydrates and

proteins (Chen et al., 2014), and the thermal degradation peaks of carbohydrates and proteins

normally merge together. The peak of S. obliquus CNW-N accompanied by a shoulder or those

of C. vulgaris ESP-31, and C. sp. JSC4 accompanied by smaller peaks are the consequence of

thermal degradation of lipids. This arises from the fact that the pyrolytic temperatures of lipids

are higher than those of carbohydrates and proteins (López-González et al., 2015).

From TGA and DTG curves, overall, the microalgae pyrolysis can be divided into three

different reactive stages, which are separated by vertical lines shown in Figure 1. In the first

stage, which lasts to 200 °C, water is removed from microalgae. The mass loss in this stage

depends strongly on the moisture content of feedstock. The second stage, from 200 to 600 °C, is

the most reactive and accounts for the majority of the mass loss during pyrolysis. In this stage,

all microalgal components (carbohydrates, proteins, lipids, and other minor components) are

decomposed to produce chars and release volatiles. The third stage starts from 600 °C, which

shows slight mass loss due to the degradation of carbonaceous matters in the solid residues

(Rizzo et al., 2013).

3. Pyrolysis of microalgae

3.1. Pyrolysis of microalgae and classification

Pyrolysis is a flexible way to produce bio-oil and biochar for sustainable or green fuel

production. The production of bio-oil and biochar depend on the pyrolysis process or reactor

adopted. Non-condensable gases, containing H2, CO, CO2, CH4, C2H2, C2H6, and other gaseous

hydrocarbons, are also produced. Microalgae pyrolysis can be classified into four modes: (1)

slow pyrolysis, (2) fast pyrolysis, (3) catalytic pyrolysis, and (4) microwave pyrolysis (Chen et

al., 2015a) where three different reactors of fixed bed, fluidized bed, and auger reactors can be

8
adopted. In general, fixed bed reactors are used in slow and microwave pyrolysis, and fluidized

bed reactors are utilized in fast pyrolysis. In regard to auger reactors, they have also been

applied for catalytic pyrolysis. A number of studies of microalgae pyrolysis are tabulated in

Table 2 in which the four types of pyrolysis, reactors, operating conditions, and bio-oil and

biochar yields are summarized.

In slow pyrolysis, biomass is converted into bio-oil and biochar at temperatures of

350-600 °C along with a low heating rate ( ≤ 10 °C min-1 ) and a long residence time of hot

vapor (10-30 s) (Miao et al., 2004). As can be seen in Table 2, the bio-oil and biochar yields

from the slow pyrolysis of microalgae are in the ranges of 18-55 wt% and 10-40 wt%,

respectively. It was reported (Demirbaş, 2006) that the bio-oil yield from the slow pyrolysis of

microalgae increased with temperature until approximately 500 °C. Thereafter, the yield

decreased when the temperature was further increased. In contrast, fast pyrolysis is featured by

a high heating rate (3,000-36,000 °C min-1) and a short residence time of hot vapor (1-3 s)

(Chen et al., 2015a), and the pyrolysis temperature is normally operated between 450 and 700

°C. To achieve the high heating rate for high bio-oil yield, small biomass particles are normally

fed into the fluidized bed pyrolyzers. The bio-oil and biochar yields from the fast pyrolysis of

microalgae are in the ranges of 18-72 wt% and 22-63 wt%, respectively.

Microalgae pyrolyzed under the aid of catalyst is termed catalytic pyrolysis where Na2CO3,

MgO, ZnCl2, and ZSM-5-based zeolites (e.g. H-ZSM-5, Fe-ZSM-5 Cu-ZSM-5 and Ni-ZSM-5)

can be utilized as the catalysts (Babich et al., 2011; Campanella & Harold, 2012). The

advantages using catalytic pyrolysis for microalgae are that bio-oil with less oxygenic

compounds, lower acidity, and higher aromatics, calorific value, and bio-oil yield can be

obtained (Suali & Sarbatly, 2012), and catalysts used for the pyrolysis can be recycled into the

reactor (Babich et al., 2011). Catalytic pyrolysis is usually operated at temperatures of 300-600

9
°C and catalyst-to-biomass mass ratios of 0.2-5 (Chen et al., 2015a). The bio-oil and biochar

yields produced from catalytic pyrolysis are approximately in the ranges of 20-58 wt% and

16-48 wt% (Babich et al., 2011; Pan et al., 2010). A schematic of the aforementioned

microalgae pyrolysis methods and bio-oil and biochar yields is shown in Figure 2.

Dielectric materials are able to convert microwaves into thermal energy through a

dielectric heating process (Chen & Lin, 2010). Accordingly, bio-oil and biochar can be

produced from microalgae through microwave pyrolysis along with microwave absorbers (e.g.

SiC, Fe3O4, CuO, water, fats, activated carbon, biochar, ionic liquids, sulfuric acid, etc).

Microwave-assisted heating possesses a number of merits over traditional heating such as rapid

heating, uniform internal heating of feedstock, instantaneous response for rapid start-up and

shut down, no need for agitation via fluidization, and hence fewer particles (ashes) in the

produced bio-oil (Borges et al., 2014; Chen et al., 2015a). Microwave pyrolysis is usually

operated at temperatures of 450-800 °C and absorber contents of 5-30 wt%. The bio-oil and

biochar yields from microwave pyrolysis are in the ranges of 21-59 wt% and 14-49 wt%,

respectively (Borges et al., 2014; Chen et al., 2015a).

Among the pyrolysis products, biochar production from slow pyrolysis is a mature

technology; while bio-oil production are still being developed, employing several advanced

technologies such as fast, catalytic pyrolysis, and microwave pyrolysis as introduced earlier.

Also, practical applications of biochar are wider than those of bio-oil, which will be discussed

in the next subsections.

3.2. Biochar production

Biochar produced from biomass can be applied in agriculture and water treatment (Xu et

al., 2012); it can also be consumed as fuel and reducing agents (Chen et al., 2015b). From

energy point of view, it has been reported that bio-oil and biochar represented 57% and 36% of

10
energy contents of microalgae feedstock, respectively (Wang et al., 2013). In particular, biochar

offers numerous benefits when applied to soils. Most of the volatiles in biomass have already

been driven off after undergoing pyrolysis, and the residual biochar is highly stable chemically

and biologically. On account of rich carbon contained in biochar which can remain stable in soil

for hundred or even thousand years (Chaiwong et al., 2013), biochar has been thought of as a

potential carbon sequestration agent (Laird et al., 2009). This implies, in turn, that there exists

an opportunity in establishing a biological carbon capture and storage solution through biomass

pyrolysis to produce biochar or long-term carbon sink, using soil as the storage media (Grierson

et al., 2011). This substantially delays the release of CO2 into the atmosphere and thereby

mitigates global warming. Biochar contains most of the feedstock’s mineral components so that

it can be used as a soil amendment, and the high inorganic contents (e.g. potassium,

phosphorous, and nitrogen) in biochar may be suitable to provide nutrients for crop production

(Wang et al., 2013). Moreover, biochar has a highly porous structure. The addition of bio-char

into soil could improve water retention and increase the surface area of the soil. This facilitates

the efficiency of nutrient use (Chaiwong et al., 2013).

3.3. Bio-oil production

Bio-oil produced from microalgae pyrolysis has several environmental advantages over

fossil oil. For example, bio-oil is CO2 or greenhouse gas (GHG) neutral so that it generates

carbon dioxide credits. Microalgae contain extremely low amount of sulfur (Kim et al., 2014);

hence there are almost no SOx emissions when microalgae-based bio-oil is burned. Meanwhile,

it was addressed that pyrolytic bio-oil from microalgae had lower oxygen and water contents

and higher calorific value and carbon content than that from lignocellulosic biomass (Rizzo et

al., 2013). However, it is notable that bio-oil from microalgae showed higher nitrogen content

compare with petroleum oils, likely due to nitrogen contained in proteins inside the microalgae.

11
This leads to higher NOx emissions in fuel use and catalyst poisoning in deoxygenation process

for further processing (Wildschut et al., 2010). Due to the inteferior properties compared to

petroleum oil, the applications of pyrolysis bio-oil is still limited now. Nevertheless, bio-oil can

be used not only as fuel but also as raw material for chemical production. According to (Xiu &

Shahbazi, 2012), industrial applications of bio-oil include:

 Combustion fuel for heat and power generation.

 Transportation fuel after upgrading.

 Production of pharmaceuticals, surfactants, and biodegradable polymers.

 Liquid smoke and wood flavors.

 Production of adhesives, chemicals, and resins.

4. Thermogravimetric analysis (TGA) technique for pyrolysis kinetics of microalgae

4.1. Principle of TGA

TGA pertains to a thermal analysis technique in nature in which the weight of a sample is

continuously measured under a controlled temperature program. Based on the variation of

sample weight with respect to temperature or time, the differential thermogravimetric (DTG)

curve can be further obtained by differentiate the TGA curve where the physical and chemical

properties of samples as a function of temperature can be determined (Yahiaoui et al., 2015).

The utilization of thermogravimetric analyzer has the advantages of easy operation, minimal

quantity of feedstock, precise control and record of temperature and sample weight loss. TGA

has been used to perform the proximate analysis of biomass (Saldarriaga et al., 2015) and figure

out the thermal behavior of microalgae such as combustion (Tang et al., 2011), ignition and

burnout (Chen et al., 2016), torrefaction (Chen et al., 2014), pyrolysis (Sanchez-Silva et al.,

2013), gasification (Sanchez-Silva et al., 2013), and chemical kinetics (Bach & Chen, 2017).

12
The relevant studies using TGA as a tool are shown in Table 3.

TGA has received immense attention in understanding solid fuel degradation to release

energy. TGA is able to trace the thermal degradation patterns of fresh biomass, biochar, and

hydrochar under air or inert environment for subsequent kinetic studies (Bach et al., 2017b;

Broström et al., 2012). The patterns reveal the response of heating rate, the inherent biomass

properties and severity that define the char combustion profile in the form of thermograph

(Islam et al., 2016).

4.2. Fundamental kinetic expressions

It is known that the knowledge of thermal behavior and pyrolysis kinetics of microalgae is

very crucial for recognizing the thermal degradation (or stability) and conversion of substance

and product formation; it is also conducive to process rate prediction, experimental design,

operational parameter decision, and practical operation. In general, the pyrolysis kinetics is

expressed in terms of the Arrhenius law which comprises activation energy, frequency factor,

and reaction order. The reaction rate under an isothermal condition as a function of time is

described as:

(1)

where is the conversion degree, is the conversion time, is the pre-exponential factor,

is the activation energy of the reaction, is the universal gas constant, and is the absolute

temperature. The conversion degree ( ) is defined as the mass fraction of decomposed solid or

released volatiles:

(2)

where and are the initial and final masses of solid, is the mass of solid at any time;

is the total mass of released volatiles, and is the mass of released volatiles at any time.

13
The reaction rate in Eq. (1) can be easily transformed into a non-isothermal expression,

which describes the conversion rate as a function of temperature at a constant heating rate, , as

the following:

(3)

By substituting Eq. (1) into Eq. (3), the non-isothermal rate expression is obtained as:

(4)

The function in Eqs. (1) and (4) can be expressed by several equations depending on

the selection of reaction mechanisms. A list of expressions for is adopted from (White et

al., 2011) and presented in Table 4. Among those, mechanisms based on reaction order are

commonly used for most biomass pyrolysis.

4.3. Kinetic models

This part reviews common pyrolysis kinetic models that have been successfully applied

for microalgae. Like other biomass materials, microalgae pyrolysis virtually consists of

numerous chemical reactions, and a number of intermediates and substances are produced

during the process. In the simplest kinetics approach, pyrolysis activation energy and

pre-exponential factor can be approximately calculated from TGA data without any knowledge

about the reaction mechanism. These approximations are called “kinetic-free model” because

they offer very limited kinetic information, except the activation energy and pre-exponential

factor. On the other hand, more comprehensive kinetic approaches propose that microalgae

produce chars and volatiles during pyrolysis, regardless of the actual number of products (Di

Blasi, 2008). Based on this idea, one or multiple reaction models are proposed, in which

microalgal components are directly converted to char products in only one stage. In another

pyrolysis kinetic model, it assumes that microalgae form intermediates and initial volatiles in

14
the primary reaction. Then, the intermediates are converted into final chars and additional

volatiles. The latter model is known as the two-step or consecutive-reaction model. Volatiles in

these models include both permanent and condensable gases at high temperatures. Because

condensable gases become liquid when they are cooled, the formation of liquid products is not

considered in these models.

4.3.1. Kinetic-free models

The objective of this method is to calculate the activation energy and pre-exponential

factor from experimental TGA data at any conversion rates through linear transformations of

Eq. (4). By omitting reaction mechanism, Eq. (1) can be rearranged as:

(5)

Integrating both sides of Eq. (5) leads to the following equation:

(6)

Different mathematical methods including transformations, approximations, and

simplifications can be applied to convert Eq. (6) into linear equations, e.g.

Kissinger–Akahira–Sunose (KAS) method in Eq. (7) and Flynn–Wall–Ozawa (FWO) method

in Eq. (8) (Shuping et al., 2010).

(7)

(8)

At any conversion rate ( ), the plot of or versus inverted temperature

should create straight lines, from which the activation energy ( ) and the pre-exponential

factor ( ) can be calculated through the slope and the intercept, respectively. Generally, the

15
conversion rates vary from 0.1 to 0.9 with an interval of 0.1. Plotting of these equations is

presented in Figure 3 to demonstrate the FWO and KAS methods (Tran et al., 2014). .

4.3.2. Single reaction model

In this model, all components in a microalga are supposed to have the same thermal

reactivity, thus only one reaction, as shown below, is needed to elucidate the microalga

pyrolysis.

(9)

Despite the simplicity of this model, it normally offers very poor fit quality due to the

different reactivities of the components in the microalga.

4.3.3. Multiple parallel reaction models

A general form of multiple parallel reaction models is presented in Eq. (10). These models

are bases on a three parallel reaction model, which has been successfully applied for

lignocellulosic biomass for decades (Di Blasi, 2008; Orfão et al., 1999). Due to difference in the

composition between lignocellulosic and microalgal biomass species, the number of parallel

reactions was modified to be achieve the best fit quality for modeling microalgal pyrolysis

(Bach & Chen, 2017; Bui et al., 2016; Sharara et al., 2014). Hence, different assumptions for

the number of microalgal components will lead to different numbers of involved reactions in

the models.

(10)

If the microalgal components are divided into two groups based on their thermal stability,

16
a two-reaction model can be expressed as:

(11)

where LTSC and HTSC designate low thermal stable components and high thermal stable

components in microalgae, respectively (Chen et al., 2015a). Considering the thermal stability

of microalgal components, the former presents carbohydrates and proteins, while the latter

stands for lipids (Bach & Chen, 2017).

A microalga generally consists of three main components (carbohydrate, protein, and lipid)

and other minor components. Due to different thermal reactivities, these components behave

differently during pyrolysis. Thus, appropriate models should include three (excluding other

components) or four (including other components) parallel reactions to better describe the

pyrolysis of microalga, as shown in Eq. (12).

(12)

Furthermore, a more detailed model may contain up to seven parallel reactions. In this

model, carbohydrate and protein in a microalga are partitioned into two groups: high and low

thermal resistant substances. Another new component in this complex model is the intermediate

products, which are formed during the pyrolysis of a microalga at low temperatures and react

independently with other components.

17
4.3.4. Series reaction or consecutive-reaction model

Different from the aforementioned parallel reaction models, this model consists of

reactions in series: initial microalga is converted to Intermediate and Volatile 1 in Eq. (13);

thereafter, the Intermediate forms final Char and Volatile 2 in Eq. (14). This series reaction

model was most applied for the isothermal kinetic study (Bach et al., 2016; Bates & Ghoniem,

2012).

(13)

(14)

It should be kept in mind that the conversion of the intermediate in the secondary reaction

depends on the degradation of initial microalga in the primary reaction. This assumption is

different from the aforementioned seven parallel reaction model, where intermediate products

react independently with other components. (Branca & Di Blasi, 2004) compared the parallel

and series reaction models for lignocellulose biomass and found that kinetic parameters

obtained in consecutive combustion were invariant with the selection of model. However, no

similar work for microalgae has been reported.

4.3.5. Distributed activation energy model

All the aforementioned models assume that the activation energy values of the microalgal

components are constant but a component may consist of several reacting species whose

reactivities are different. These variations can lead to the fact that the activation energy of a

component is not a constant value but in a range of values ( rhe i et al., 2010). On

accounting for this issue, a distributed activation energy model (DAEM) can employed for

pyrolysis kinetic study. In the DAEM, complex pyrolysis decomposition of any microalagal

component can be described by a series of reactions having different activation energy values

18
but the same pre-exponential factor. A general equation for the DAEM can be written as:

(15)

The is the distribution function of activation energy, which can be described by

Gaussian, Weibull, or Gamma distribution (de Caprariis et al., 2012). Among those, the

Gaussian function (Eq 16), which show a mean value (E0) and its deviation σ, is used more

commonly.

(16)

Although the DAEM offers much better fit quality than other models, the model has a

double-layer integral and one variable ( ) goes from zero to infinite (Eq. 15), which can not be

calculated directly. However, some simplifications (Miura, 1995; Please et al., 2003) were

proposed to reduce the complexity of the model in practical calculation.

4.4. Progress in pyrolysis kinetics study for microalgae

Currently, the number of kinetic studies on microalgae pyrolysis via TGA is much smaller

than that of pyrolysis studies for bio-oil and biochar production in other reactors (e.g. fixed bed,

fluidized bed, auger, etc.). An up-to-date list of publications for TGA pyrolysis studies are

presented in Table 5. From the table, typical ranges for TGA operating conditions can be

established. TGA pyrolysis normally requires amounts of 5-20 mg microalga and inert gas

(nitrogen or argon) flow rates of 50-200 mL/min. Moreover, heating rates of 5-40 °C/min can

be applied to reach final temperatures at 700-1000 °C. Because all works in Table 5 employed

non-isothermal mode for TGA pyrolysis in order to ensure completed pyrolysis, the final

temperatures in these studies are virtually higher than those for pyrolysis in other reactors

(Table 2), which were carried out under an isothermal mode at the final temperatures or at very

19
high heating rates.

Although various methods has been applied to study pyrolysis kinetics of microalgae,

most of the works listed in Table 5 employed kinetic-free models, presented in section 4.3.1, to

calculate a range of pyrolysis activation energy and pre-exponential factor at various

conversion rates (normally, from 0.1 to 0.9 with intervals of 0.1). (Shuping et al., 2010)

reported that pyrolysis activation energy of microalga Dunaliella tertiolecta did not show large

variations when calculated by Kissinger–Akahira–Sunose (KAS) and Flynn–Wall–Ozawa

(FWO) methods. The activation energy values were 131.7-152.7 kJ/mol by KAS method and

134.4-152.9 kJ/mol by FWO method, respectively. Similarly, (Agrawal & Chakraborty, 2013)

found that the activation energy of Chlorella vulgaris were 43.7-67.6 kJ/mol by FWO method

and 41.2-63.7 kJ/mol by KAS method. By adopting Freeman–Caroll method, researchers

revealed that the activation energy of Chlorella spp. were 71.3-79.2 kJ/mol (Rizzo et al., 2013),

and those of S. Platensis and C. Potothecoides were 76–97 kJ/mol and 42–52 kJ/mol (Peng et

al., 2001). In addition, (Gai et al., 2013) employed Vyazovkin method and showed that

activation energy of Chlorella pyrenoidosa and Spirulina platensis were 8.9-114.5 kJ/mol and

74.4-140.1 kJ/mol, respectively. Due to limited investigation points, these models are unable to

reproduce thermogravimetric curves or provide any information about the fit quality between

the modeled and experimental TGA data for model evaluation. Few studies divided DTG

curves into different zones (Agrawal & Chakraborty, 2013) or decreased conversion rate

intervals (down to 0.01) (Hu et al., 2015) to increase the number of investigation points, which

can approximately draw predicted curves. In other study, (Sanchez-Silva et al., 2013) employed

regression analysis to interpolate additional points and provided statistical R2 values to evaluate

their model. Though some improvements and modifications have been applied for kinetic-free

models to help create modeled curves and provide fit quality information, these models cannot

20
provide the reaction mechanism, e.g. how microalgal components behave during pyrolysis. To

overcome this issue, multiple parallel reaction models (section 4.3.3) are more appropriate. The

number of independent reactions can be four (Sharara et al., 2014), five (Bui et al., 2016), or up

to seven (Bach & Chen, 2017). Moreover, each reaction represents the degradation of each

microalgal component. (Sharara et al., 2014) assumed that component 1 includes moisture and

light hydrocarbons, component 2 and 3 are respectively protein and starch, while component 4

represent passive pyrolysis weight loss. On the other hand, (Bui et al., 2016) assigned five

components as hemicellulose, cellulose, lignin, lipid, and protein. Recently, (Bach & Chen,

2017) examined different one-step models and changed the number of involved reactions, from

one to seven. The identical names of the components in these models are summarized in Table 6.

They compared the fit quality of the models with increasing the number of reactions and found

that the more reactions the model has, the better fit quality it offers. Improvement in fit quality

with increasing the number of reactions is visually demonstrated in Figure 4. It can be seen that

the unique advantage of the multiple reaction models compared with the kinetic-free model is

that the former can provide the information about the reactivity of each component in a

microalga.

Recenty, the series reaction model (section 4.3.4) has not been applied for microalga

kinetic study yet. It may be due to the fact that this model is more suitable to investigate TGA

data in isothermal mode (Bach et al., 2016; Bates & Ghoniem, 2012), rather than common

non-isothermal TGA employed in most recent pyrolysis studies. Although it can be found that

(Sanchez-Silva et al., 2013) employed series reactions to investigate the release of gases during

pyrolysis, it requires additional mass spectrometry to record the evolution of produced gases

and not really employs TGA technique.

Compared to multiple parallel reaction models, application of the DAEM for microalgal

21
pyrolysis is less common. Although the DAEM is widely applied for kinetic studies of several

lignocellulosic biomass species, only a few works (Ceylan & Kazan, 2015; Hu et al., 2015)

employed simplified DAEM for pyrolysis kinetics of microalgae N. oculata, Tetraselmis sp.,

Chlorella pyrenoidosa and bloom-forming cyanobacteria. However, very little information

about the advantages of the DAEM can be ontained from these studies despite that the authors

concluded that the DAEM was better than a single-step model and it gave excellent fits between

simulated results and experimental data. In addition, Hu (2015) found that activation energy

values estimated from the DAEM were always lower than those from the single-step global

model. In detail, the activation energy of Chlorella pyrenoidosa and bloom-forming

cyanobacteria from the DAEM were respectively 100.6 and 144.5 kJ/mol, while these values

from the single-step global model were 143.71 and 173.46 kJ/mol.

4.5. Potential for process design and up-scaling using kinetic data

Seeing that pyrolysis is an important process for producing bio-oil and biochar from

microalgae, it has received growing attention recently. Pyrolysis kinetics study thus plays an

important role in the design and optimization of pyrolyzers and it is worth to investigate the

kinetics prior to practical applications (Doyle, 1961). According to (Hakvoort et al., 1989), the

thermal conversion curves may differ for identical experiment conditions but the average

activation energy should be nearly constant despite different heating rates and heating programs.

Therefore, TGA becomes the most favorable tool for kinetics studies of biomass materials

including microalgae for several decades (Várhegyi, 2007), even though heating profile in TGA

experiment is far from that for practical situations.

Currently, pyrolysis of microalgae has not been exploited at industrial scale yet. However,

kinetic data from the available studies could be benefical for the process design and up-scaling.

Prior to industrial deployment, it needs to study conceptual process modeling and optimization

22
using commercial simulators (e.g., Aspen Plus). Modeling of pyrolysis can employ kinetic

reaction models based on thermodynamic equilibrium calculations, which can offer flexible

and predictive simulations for a wide range of biomass feedstock (Peters et al., 2017). Moreover,

these kinetic schemes use power law kinetic expressions, which is quite similar to the

fundamental Arrhenius’ equation and has been widel used in p rol sis kinetic studies. Results

obtained from these models can provide an overview on the pyrolysis processes at plant level as

well as assessments on energy, economic and even environmental impacts of the process. In

short, kinetic study is always the first step and essential to provide information for further

process design and up-scaling.

5. Challenges and opportunities

Due to the limited number of publications reporting on multiple parallel reaction models, it

suggests that further kinetic studies would employ these models due to their advantages.

Among the studies, a study (Bach & Chen, 2017) provided comprehensive kinetic information

but only limited to Chlorella vulgaris ESP-31, applications of the models for other species are

strongly recommended. Moreover, some conflictions are found among the studies when

assigning reactions to components. Addressing this issue in further works is also advisable.

Last but not least, application of the DAEM for microalgal pyrolysis kinetic study needs more

attention due to the excellent fit of the DAEM compared to other models.

However, it should be reminded that this technique aims at investigating only the intrinsic

(or chemical) kinetics, i.e. heat and mass transfer phenomena are excluded. Thus, a small

amount of microalgal sample with fine particles and not too high temperatures must be

employed to establish a chemical kinetic control regime. Depending on feedstock

characteristics, different sample weights can be found in the literature. However, an amount of

23
5 mg or less with particles sizes less than 150 µm are recommended for TGA pyrolysis study. In

addition, a final temperature in the range of 700-800 °C can be employed for non-isothermal

mode. In practical applications, complex interactions between the chemistry and transport

phenomena always occur at inter- and intra-particles. Thus it recommends exploring more

detailed models coupling with heat, mass and momentum transports to further understand the

behaviors of microalgal particles during practical pyrolysis.

6. Conclusions

A comprehensive overview on recent research activities in TGA pyrolysis characteristics

and kinetics of various microalgae has been presented, which is conducive to pyrolyzer design,

operation optimization, and biofuel production. For kinetics, the kinetic-free models are simple,

but provide very limited kinetic information. Therefore, multiple parallel reaction models are

more preferable for further intrinsic kinetic studies, which are able to clarify the reactions of

individual microalgal components. Last but not least, more attention on the DAEM and detailed

models coupling with heat, mass and momentum transports to further understand the pyrolysis

behaviors of microalgal particles in practical situations are recommended.

Acknowledgments

The authors acknowledge the financial support from the Ministry of Science and

Technology, Taiwan, R.O.C., under the contract MOST 106-2923-E-006-002-MY3 for this

research. The authors also thank Prof. Jo-Shu Chang at National Cheng Kung University for

providing the microalgae in this study.

References

1. Agrawal, A., Chakraborty, S. 2013. A kinetic study of pyrolysis and combustion of


microalgae Chlorella vulgaris using thermo-gravimetric analysis. Bioresource Technology,
128, 72-80.

24
2. Babich, I. ., van der Hulst, M., Lefferts, L., Moulijn, J.A., O’Connor, P., Seshan, K. 2011.
Catalytic pyrolysis of microalgae to high-quality liquid bio-fuels. Biomass and Bioenergy,
35(7), 3199-3207.
3. Bach, Q.-V., Chen, W.-H. 2017. A comprehensive study on pyrolysis kinetics of
microalgal biomass. Energy Conversion and Management, 131, 109-116.
4. Bach, Q.-V., Chen, W.-H., Chu, Y.-S., Skreiberg, Ø. 2016. Predictions of biochar yield and
elemental composition during torrefaction of forest residues. Bioresource technology, 215,
239-246.
5. 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.
6. Bach, Q.-V., Tran, K.-Q., Skreiberg, Ø. 2017b. Comparative study on the thermal
degradation of dry- and wet-torrefied woods. Applied Energy, 185, Part 2, 1051-1058.
7. Bates, R.B., Ghoniem, A.F. 2012. Biomass torrefaction: Modeling of volatile and solid
product evolution kinetics. Bioresource Technology, 124, 460-469.
8. Borges, F.C., Xie, Q., Min, M., Muniz, L.A.R., Farenzena, M., Trierweiler, J.O., Chen, P.,
Ruan, R. 2014. Fast microwave-assisted pyrolysis of microalgae using microwave
absorbent and HZSM-5 catalyst. Bioresource Technology, 166, 518-526.
9. Branca, C., Di Blasi, C. 2004. Parallel- and series-reaction mechanisms of wood and char
combustion. Thermal science, 8, 51-63.
10. Brennan, L., Owende, P. 2010. Biofuels from microalgae—A review of technologies for
production, processing, and extractions of biofuels and co-products. Renewable and
Sustainable Energy Reviews, 14(2), 557-577.
11. Broström, M., Nordin, A., Pommer, L., Branca, C., Di Blasi, C. 2012. Influence of
torrefaction on the devolatilization and oxidation kinetics of wood. Journal of Analytical
and Applied Pyrolysis, 96(0), 100-109.
12. Bui, H.-H., Tran, K.-Q., Chen, W.-H. 2016. Pyrolysis of microalgae residues – A kinetic
study. Bioresource Technology, 199, 362-366.
13. Campanella, A., Harold, M.P. 2012. Fast pyrolysis of microalgae in a falling solids reactor:
Effects of process variables and zeolite catalysts. Biomass and Bioenergy, 46, 218-232.
14. Ceylan, S., Kazan, D. 2015. Pyrolysis kinetics and thermal characteristics of microalgae
Nannochloropsis oculata and Tetraselmis sp. Bioresource Technology, 187, 1-5.
15. 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.
16. Chen, C., Ma, X., He, Y. 2012a. Co-pyrolysis characteristics of microalgae Chlorella
vulgaris and coal through TGA. Bioresource Technology, 117, 264-273.
17. Chen, W.-H., Lin, B.-J. 2010. Effect of microwave double absorption on hydrogen
generation from methanol steam reforming. International Journal of Hydrogen Energy,
35(5), 1987-1997.
18. Chen, W.-H., Lin, B.-J., Huang, M.-Y., Chang, J.-S. 2015a. Thermochemical conversion
of microalgal biomass into biofuels: a review. Bioresource technology, 184, 314-327.
19. Chen, W.-H., Lu, K.-M., Tsai, C.-M. 2012b. An experimental analysis on property and
structure variations of agricultural wastes undergoing torrefaction. Applied energy, 100,
318-325.
20. Chen, W.-H., Peng, J., Bi, X.T. 2015b. A state-of-the-art review of biomass torrefaction,
densification and applications. Renewable and Sustainable Energy Reviews, 44, 847-866.

25
21. Chen, W.-H., Wu, Z.-Y., Chang, J.-S. 2014. Isothermal and non-isothermal torrefaction
characteristics and kinetics of microalga Scenedesmus obliquus CNW-N. Bioresource
Technology, 155, 245-251.
22. 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.
23. Chiaramonti, D., Prussi, M., Buffi, M., Rizzo, A.M., Pari, L. 2017. Review and
experimental study on pyrolysis and hydrothermal liquefaction of microalgae for biofuel
production. Applied Energy, 185, Part 2, 963-972.
24. Chiu, S.-Y., Kao, C.-Y., Chen, T.-Y., Chang, Y.-B., Kuo, C.-M., Lin, C.-S. 2015.
Cultivation of microalgal Chlorella for biomass and lipid production using wastewater as
nutrient resource. Bioresource technology, 184, 179-189.
25. Cho, S.-H., Kim, K.-H., Jeon, Y.J., Kwon, E.E. 2015. Pyrolysis of microalgal biomass in
carbon dioxide environment. Bioresource Technology, 193, 185-191.
26. de Caprariis, B., De Filippis, P., Herce, C., Verdone, N. 2012. Double-Gaussian
Distributed Activation Energy Model for Coal Devolatilization. Energy & Fuels, 26(10),
6153-6159.
27. Demirbaş, A. 2006. Oil Products from Mosses and Al ae via P rol sis. Energy Sources,
Part A: Recovery, Utilization, and Environmental Effects, 28(10), 933-940.
28. Di Blasi, C. 2008. Modeling chemical and physical processes of wood and biomass
pyrolysis. Progress in Energy and Combustion Science, 34(1), 47-90.
29. Dong, T., Gao, D., Miao, C., Yu, X., Degan, C., Garcia-Pérez, M., Rasco, B., Sablani, S.S.,
Chen, S. 2015. Two-step microalgal biodiesel production using acidic catalyst generated
from pyrolysis-derived bio-char. Energy Conversion and Management, 105, 1389-1396.
30. Doyle, C.D. 1961. Kinetic analysis of thermogravimetric data. Journal of Applied Polymer
Science, 5(15), 285-292.
31. Du, S.-W., Chen, W.-H. 2006. Numerical prediction and practical improvement of
pulverized coal combustion in blast furnace. International Communications in Heat and
Mass Transfer, 33(3), 327-334.
32. Francavilla, M., Kamaterou, P., Intini, S., Monteleone, M., Zabaniotou, A. 2015.
Cascading microalgae biorefinery: Fast pyrolysis of Dunaliella tertiolecta lipid
extracted-residue. Algal Research, 11, 184-193.
33. Gai, C., Zhang, Y., Chen, W.-T., Zhang, P., Dong, Y. 2013. Thermogravimetric and kinetic
analysis of thermal decomposition characteristics of low-lipid microalgae. Bioresource
Technology, 150, 139-148.
34. Glaser, B., Parr, M., Braun, C., Kopolo, G. 2009. Biochar is carbon negative. Nature
Geosci, 2(1), 2-2.
35. Grierson, S., Strezov, V., Ellem, G., McGregor, R., Herbertson, J. 2009. Thermal
characterisation of microalgae under slow pyrolysis conditions. Journal of Analytical and
Applied Pyrolysis, 85(1–2), 118-123.
36. Grierson, S., Strezov, V., Shah, P. 2011. Properties of oil and char derived from slow
pyrolysis of Tetraselmis chui. Bioresource Technology, 102(17), 8232-8240.
37. Hakvoort, G., Schouten, J.C., Valkenburg, P.J.M. 1989. The determination of coal
combustion kinetics with thermogravimetry. Journal of Thermal Analysis and
Calorimetry, 35(2), 335-346.

26
38. Hu, M., Chen, Z., Guo, D., Liu, C., Xiao, B., Hu, Z., Liu, S. 2015. Thermogravimetric
study on pyrolysis kinetics of Chlorella pyrenoidosa and bloom-forming cyanobacteria.
Bioresource Technology, 177, 41-50.
39. Hu, Z., Ma, X., Li, L. 2016. The synergistic effect of co-pyrolysis of oil shale and
microalgae to produce syngas. Journal of the Energy Institute, 89(3), 447-455.
40. Hu, Z., Ma, X., Li, L., Wu, J. 2014. The catalytic pyrolysis of microalgae to produce
syngas. Energy Conversion and Management, 85, 545-550.
41. Islam, M.A., Auta, M., Kabir, G., Hameed, B.H. 2016. A thermogravimetric analysis of
the combustion kinetics of karanja (Pongamia pinnata) fruit hulls char. Bioresource
Technology, 200, 335-341.
42. Janković, ., Schultz, D.M. 2017. Atmosfear: Communicatin the Effects of Climate
Change on Extreme Weather. Weather, Climate, and Society, 9(1), 27-37.
43. Jena, U., Das, K.C. 2011. Comparative Evaluation of Thermochemical Liquefaction and
Pyrolysis for Bio-Oil Production from Microalgae. Energy & Fuels, 25(11), 5472-5482.
44. Kebelmann, K., Hornung, A., Karsten, U., Griffiths, G. 2013. Intermediate pyrolysis and
product identification by TGA and Py-GC/MS of green microalgae and their extracted
protein and lipid components. Biomass and Bioenergy, 49, 38-48.
45. Kim, S.-S., Ly, H.V., Choi, G.-H., Kim, J., Woo, H.C. 2012. Pyrolysis characteristics and
kinetics of the alga Saccharina japonica. Bioresource Technology, 123, 445-451.
46. Kim, S.-S., Ly, H.V., Kim, J., Choi, J.H., Woo, H.C. 2013. Thermogravimetric
characteristics and pyrolysis kinetics of Alga Sagarssum sp. biomass. Bioresource
Technology, 139(0), 242-248.
47. Kim, S.W., Koo, B.S., Lee, D.H. 2014. A comparative study of bio-oils from pyrolysis of
microalgae and oil seed waste in a fluidized bed. Bioresource Technology, 162, 96-102.
48. Kirtania, K., Bhattacharya, S. 2012a. Application of the distributed activation energy
model to the kinetic study of pyrolysis of the fresh water algae Chlorococcum humicola.
Bioresource Technology, 107, 476-481.
49. Kirtania, K., Bhattacharya, S. 2012b. Application of the distributed activation energy
model to the kinetic study of pyrolysis of the fresh water algae Chlorococcum humicola.
Bioresource Technology, 107(0), 476-481.
50. Lü, J., Sheahan, C., Fu, P. 2011. Metabolic engineering of algae for fourth generation
biofuels production. Energy & Environmental Science, 4(7), 2451-2466.
51. López-González, D., Fernandez-Lopez, M., Valverde, J.L., Sanchez-Silva, L. 2014.
Pyrolysis of three different types of microalgae: Kinetic and evolved gas analysis. Energy,
73, 33-43.
52. López-González, D., Puig-Gamero, M., Acién, F.G., García-Cuadra, F., Valverde, J.L.,
Sanchez-Silva, L. 2015. Energetic, economic and environmental assessment of the
pyrolysis and combustion of microalgae and their oils. Renewable and Sustainable Energy
Reviews, 51, 1752-1770.
53. Laird, D.A., Brown, R.C., Amonette, J.E., Lehmann, J. 2009. Review of the pyrolysis
platform for coproducing bio‐ oil and biochar. Biofuels, Bioproducts and Biorefining, 3(5),
547-562.
54. Lee, J.W., Hawkins, B., Day, D.M., Reicosky, D.C. 2010. Sustainability: the capacity of
smokeless biomass pyrolysis for energy production, global carbon capture and
sequestration. Energy & Environmental Science, 3(11), 1695-1705.
55. Li, F., Srivatsa, S.C., Batchelor, W., Bhattacharya, S. 2017. A study on growth and
pyrolysis characteristics of microalgae using Thermogravimetric Analysis-Infrared

27
Spectroscopy and synchrotron Fourier Transform Infrared Spectroscopy. Bioresource
Technology, 229, 1-10.
56. Manyà, J.J. 2012. Pyrolysis for Biochar Purposes: A Review to Establish Current
Knowledge Gaps and Research Needs. Environmental Science & Technology, 46(15),
7939-7954.
57. Marcilla, A., Gómez-Siurana, A., Gomis, C., Chápuli, E., Catalá, M.C., Valdés, F.J. 2009.
Characterization of microalgal species through TGA/FTIR analysis: Application to
nannochloropsis sp. Thermochimica Acta, 484(1–2), 41-47.
58. Mata, T.M., Martins, A.A., Caetano, N.S. 2010. Microalgae for biodiesel production and
other applications: A review. Renewable and Sustainable Energy Reviews, 14(1), 217-232.
59. Miao, X., Wu, Q. 2004. High yield bio-oil production from fast pyrolysis by metabolic
controlling of Chlorella protothecoides. Journal of Biotechnology, 110(1), 85-93.
60. Miao, X., Wu, Q., Yang, C. 2004. Fast pyrolysis of microalgae to produce renewable fuels.
Journal of analytical and applied pyrolysis, 71(2), 855-863.
61. Miura, K. 1995. A New and Simple Method to Estimate f(E) and k0(E) in the Distributed
Activation Energy Model from Three Sets of Experimental Data. Energy & Fuels, 9(2),
302-307.
62. Nakanishi, A., Aikawa, S., Ho, S.-H., Chen, C.-Y., Chang, J.-S., Hasunuma, T., Kondo, A.
2014. Development of lipid productivities under different CO 2 conditions of marine
microalgae Chlamydomonas sp. JSC4. Bioresource technology, 152, 247-252.
63. Orfão, J.J.M., Antunes, F.J.A., Figueiredo, J.L. 1999. Pyrolysis kinetics of lignocellulosic
materials—three independent reactions model. Fuel, 78(3), 349-358.
64. Pan, P., Hu, C., Yang, W., Li, Y., Dong, L., Zhu, L., Tong, D., Qing, R., Fan, Y. 2010. The
direct pyrolysis and catalytic pyrolysis of Nannochloropsis sp. residue for renewable
bio-oils. Bioresource Technology, 101(12), 4593-4599.
65. Peng, W., Wu, Q., Tu, P. 2001. Pyrolytic characteristics of heterotrophic Chlorella
protothecoides for renewable bio-fuel production. Journal of Applied Phycology, 13(1),
5-12.
66. Peters, J.F., Banks, S.W., Bridgwater, A.V., Dufour, J. 2017. A kinetic reaction model for
biomass pyrolysis processes in Aspen Plus. Applied Energy, 188, 595-603.
67. Please, C.P., McGuinness, M.J., McElwain, D.L.S. 2003. Approximations to the
distributed activation energy model for the pyrolysis of coal. Combustion and Flame,
133(1–2), 107-117.
68. Raheem, A., Azlina, W.W., Yap, Y.T., Danquah, M.K., Harun, R. 2015. Thermochemical
conversion of microalgal biomass for biofuel production. Renewable and Sustainable
Energy Reviews, 49, 990-999.
69. Razeghifard, R. 2013. Algal biofuels. Photosynthesis Research, 117(1), 207-219.
70. Rizzo, A.M., Prussi, M., Bettucci, L., Libelli, I.M., Chiaramonti, D. 2013.
Characterization of microalga Chlorella as a fuel and its thermogravimetric behavior.
Applied Energy, 102, 24-31.
71. Saldarriaga, J.F., Aguado, R., Pablos, A., Amutio, M., Olazar, M., Bilbao, J. 2015. Fast
characterization of biomass fuels by thermogravimetric analysis (TGA). Fuel, 140,
744-751.
72. Sambusiti, C., Bellucci, M., Zabaniotou, A., Beneduce, L., Monlau, F. 2015. Algae as
promising feedstocks for fermentative biohydrogen production according to a biorefinery
approach: A comprehensive review. Renewable and Sustainable Energy Reviews, 44,
20-36.

28
73. Sanchez-Silva, L., López-González, D., Garcia-Minguillan, A.M., Valverde, J.L. 2013.
Pyrolysis, combustion and gasification characteristics of Nannochloropsis gaditana
microalgae. Bioresource Technology, 130, 321-331.
74. Sharara, M.A., Holeman, N., Sadaka, S.S., Costello, T.A. 2014. Pyrolysis kinetics of algal
consortia grown using swine manure wastewater. Bioresource Technology, 169, 658-666.
75. Shuping, Z., Yulong, W., Mingde, Y., Chun, L., Junmao, T. 2010. Pyrolysis characteristics
and kinetics of the marine microalgae Dunaliella tertiolecta using thermogravimetric
analyzer. Bioresource Technology, 101(1), 359-365.
76. Singh, A., Olsen, S.I. 2011. A critical review of biochemical conversion, sustainability and
life cycle assessment of algal biofuels. Applied Energy, 88(10), 3548-3555.
77. Sirajunnisa, A.R., Surendhiran, D. 2016. Algae – A quintessential and positive resource of
bioethanol production: A comprehensive review. Renewable and Sustainable Energy
Reviews, 66, 248-267.
78. Suali, E., Sarbatly, R. 2012. Conversion of microalgae to biofuel. Renewable and
Sustainable Energy Reviews, 16(6), 4316-4342.
79. Tang, Y., Ma, X., Lai, Z. 2011. Thermogravimetric analysis of the combustion of
microalgae and microalgae blended with waste in N 2/O 2 and CO 2/O 2 atmospheres.
Bioresource technology, 102(2), 1879-1885.
80. Tran, K.-Q., Bach, Q.-V., Trinh, T.T., Seisenbaeva, G. 2014. Non-isothermal pyrolysis of
torrefied stump – A comparative kinetic evaluation. Applied Energy, 136(0), 759-766.
81. Uggetti, E., Sialve, B., Trably, E., Steyer, J.P. 2014. Integrating microalgae production
with anaerobic digestion: a biorefinery approach. Biofuels, Bioproducts and Biorefining,
8(4), 516-529.
82. Várhegyi, G. 2007. Aims and methods in non-isothermal reaction kinetics. Journal of
Analytical and Applied Pyrolysis, 79(1–2), 278-288.
83. rhe i, .b., Bob l , B.z., Jakab, E., Chen, H. 2010. Thermogravimetric Study of
Biomass Pyrolysis Kinetics. A Distributed Activation Energy Model with Prediction Tests.
Energy & Fuels, 25(1), 24-32.
84. 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.
85. Wang, X., Sheng, L., Yang, X. 2017. Pyrolysis characteristics and pathways of protein,
lipid and carbohydrate isolated from microalgae Nannochloropsis sp. Bioresource
Technology, 229, 119-125.
86. Wang, X., Zhao, B., Tang, X., Yang, X. 2015. Comparison of direct and indirect pyrolysis
of micro-algae Isochrysis. Bioresource Technology, 179, 58-62.
87. White, J.E., Catallo, W.J., Legendre, B.L. 2011. Biomass pyrolysis kinetics: A
comparative critical review with relevant agricultural residue case studies. Journal of
Analytical and Applied Pyrolysis, 91(1), 1-33.
88. Wildschut, J., Melian-Cabrera, I., Heeres, H. 2010. Catalyst studies on the hydrotreatment
of fast pyrolysis oil. Applied Catalysis B: Environmental, 99(1), 298-306.
89. Wu, K., Liu, J., Wu, Y., Chen, Y., Li, Q., Xiao, X., Yang, M. 2014. Pyrolysis characteristics
and kinetics of aquatic biomass using thermogravimetric analyzer. Bioresource
Technology, 163, 18-25.
90. Wu, X., Wu, Y., Wu, K., Chen, Y., Hu, H., Yang, M. 2015. Study on pyrolytic kinetics and
behavior: The co-pyrolysis of microalgae and polypropylene. Bioresource Technology,
192, 522-528.

29
91. Xie, Q., Addy, M., Liu, S., Zhang, B., Cheng, Y., Wan, Y., Li, Y., Liu, Y., Lin, X., Chen, P.,
Ruan, R. 2015. Fast microwave-assisted catalytic co-pyrolysis of microalgae and scum for
bio-oil production. Fuel, 160, 577-582.
92. Xiu, S., Shahbazi, A. 2012. Bio-oil production and upgrading research: A review.
Renewable and Sustainable Energy Reviews, 16(7), 4406-4414.
93. Xu, G., Lv, Y., Sun, J., Shao, H., Wei, L. 2012. Recent Advances in Biochar Applications
in Agricultural Soils: Benefits and Environmental Implications. CLEAN – Soil, Air, Water,
40(10), 1093-1098.
94. Yahiaoui, M., Hadoun, H., Toumert, I., Hassani, A. 2015. Determination of kinetic
parameters of Phlomis bovei de Noé using thermogravimetric analysis. Bioresource
technology, 196, 441-447.
95. Yang, H., Yan, R., Chen, H., Lee, D.H., Zheng, C. 2007. Characteristics of hemicellulose,
cellulose and lignin pyrolysis. Fuel, 86(12), 1781-1788.
96. Yang, X., Zhang, R., Fu, J., Geng, S., Cheng, J.J., Sun, Y. 2014. Pyrolysis kinetic and
product analysis of different microalgal biomass by distributed activation energy model
and pyrolysis–gas chromatography–mass spectrometry. Bioresource Technology, 163,
335-342.
97. Zhao, B., Wang, X., Yang, X. 2015. Co-pyrolysis characteristics of microalgae Isochrysis
and Chlorella: Kinetics, biocrude yield and interaction. Bioresource Technology, 198,
332-339.
98. Zhu, L. 2015. Biorefinery as a promising approach to promote microalgae industry: An
innovative framework. Renewable and Sustainable Energy Reviews, 41, 1376-1384.

30
Table 1. Elemetal and chemical compositions of some common microalgae.

Feedstock Elemental composition (wt%) Chemical composition (wt%) Reference

C H O N S Carbohydrate Protein Lipid Others

Dunaliella tertiolecta 39.00 5.37 53.02 1.99 0.62 21.69 61.32 2.87 n/a (Shuping et al., 2010)

Chlorella spp. 46.1 6.1 39.1 6.7 0.4 15-16.5 29.6 9-13 n/a (Rizzo et al., 2013)

Chlorella pyrenoidosa 51.2 6.8 30.7 11.3 n/a 22.5 71.5 0.2 n/a (Gai et al., 2013)

Spirulina platensis 49.6 6.2 33.4 10.8 n/a 19.3 64.7 4.8 n/a (Gai et al., 2013)

Scenedesmus almeriensis 41.9 6.7 44.7 5.9 0.8 25.2 44.2 24.6 n/a (López-González et al., 2014)

Nannochloropsis Gaditana 49.4 7.7 34.7 7.0 1.1 25.1 40.5 26.3 n/a (López-González et al., 2014)

Chlorella vulgaris 44.8 6.8 40.4 7.0 1.0 12.4 58.1 13.5 n/a (López-González et al., 2014)

Chlorella pyrenoidosa 48.56 6.80 41.29 8.39 1.76 24.07 62.42 1.83 n/a (Hu et al., 2015)

Chlorella vulgaris ESP-31 53.1 8.67 35.05 3.26 n/a 56.92 22.50 14.83 5.75 (Bach et al., 2017a)

Scenedesmus obliquus 37.37 5.80 50.02 6.82 n/a 18.77 42.53 6.52 32.17 (Chen et al., 2014)

n/a: not available

31
Table 2. A list of literature of microalgae pyrolysis.

Operating conditions Yield (wt%)


Pyrolysis
Feedstock Reactor Temp. Heating rate Sweep gas flow rate Duration Bio-oil Biochar Reference
Mode
(°C) (°C min-1) (mL min-1) (min)
Chaetocerous Slow Fixed bed 500 10 100 20 33 (Grierson et al., 2009)
muelleri
Chlorella like Slow Fixed bed 500 10 100 20 41 (Grierson et al., 2009)
Chlorella vulgaris Slow Fixed bed 500 10 100 20 41 (Grierson et al., 2009)
Spirulina platensis Slow Fixed bed 350-500 3.5-7 250 60 23-29 28-40 (Jena & Das, 2011)
Chlorella vulgaris Slow Fixed bed 530 600 (Cho et al., 2015)
Nannochloropsis Slow Fixed bed 500 180 55.21 30.26 (Wang et al., 2017)
N-heterocyclic Slow Fixed bed 475 10 400 36.8 (Wang et al., 2015)
Chlorella vulgaris Slow Fixed bed 600-800 6.67 1,333 15 11 (Hu et al., 2016)
C. protothecoides Fast Fluidized bed 500 36,000 6,667 18 (Miao et al., 2004)
C. protothecoides Fast Fluidized bed 500 36,000 6,667 58-72 (Miao & Wu, 2004)
Chlorella Fast Fluidized bed 450-550 24-43 34-63 (Chiaramonti et al.,
2017)
Dunaliella Fast Fixed bed 500-700 3,000 39-43 22-36 (Francavilla et al.,
tertiolecta 2015)
Nannochloropsis Microwave Fixed bed 550 28.6 (Xie et al., 2015)
Nannochloropsis Microwave Fixed bed 450-550 30 41-59 14-37 (Borges et al., 2014)
Chlorella sp. Microwave Fixed bed 450-550 30 41-57 26-49 (Borges et al., 2014)
Nannochloropsis Catalytic Fixed bed 300-500 10 30 120 21-31 26-46 (Pan et al., 2010)
(HZSM-5)
Chlorella vulgaris Catalytic (MgO) Fixed bed 800 1,333 8.33 20.21 16 (Hu et al., 2014)
Chlorella vulgaris Catalytic (ZnCl2) Fixed bed 800 1,333 8.33 37.11 17.77 (Hu et al., 2014)
Chlorella Catalytic (biochar Auger 600 1 47 16 (Dong et al., 2015)
sorokiniana based catalyst)
Green algae Catalytic Auger 450-600 250 38-58 25-48 (Campanella & Harold,
(H-ZSM-5) 2012)

32
Table 3. Investigation on thermal behaviors of microalgae using TGA.

Microalgae Thermal behavior Heating rate Temperature Duration Reference


(°C min-1) (°C) (min)
Chlamydomonas sp. JSC4 residue Pyrolysis (in N2) 20 105-800 (Chen et al., 2016)
Combustion, ignition and
burnout (in air)
Chlorella protothecoides Combustion 20 100-1000 (Tang et al., 2011)
C. reinhardtii Pyrolysis 100 25-900 (Kebelmann et al., 2013)
Chlorella vulgaris
Scenedesmus obliquus Pyrolysis 20 105-800 (Chen et al., 2014)
Isothermal torrefaction - 60
Non-isothermal torrefaction 0.5, 1.0, 1.5 60
Nannochloropsis sp Pyrolysis with TG-FTIR 35 50-750 (Marcilla et al., 2009)
analysis
Tetraselmis suecica Pyrolysis under different CO2 10 25-900 (Li et al., 2017)
concentrations with TG-IR
analysis
Scenedesmus sp Pyrolysis 10 125-1000 (López-González et al., 2015)
Combustion
Chlorella vulgaris Pyrolysis kinetics 10 105-700 (Bach & Chen, 2017)
Chlorella spp. Pyrolysis characteristics and 15 25-800 (Rizzo et al., 2013)
Nannochloropsis kinetics
Nannochloropsis gaditana Gasification 40 550, 650, 750, 850 60 (Sanchez-Silva et al., 2013)
Dunaliella tertiolecta and its residue Pyrolysis 10 25-800 (Francavilla et al., 2015)
D. tertiolecta Pyrolysis characteristics and 5-40 25-100 (Wu et al., 2014)
kinetics
Chlorococcum humicola Pyrolysis kinetics 5-20 25-1100 (Kirtania & Bhattacharya,
2012a)

33
Table 4. Expressions of based on common reaction mechanisms (White et al., 2011).

Reaction mechanism
Reaction order
Zero order 1
First order
nth order
Nucleation
Power law
Exponential law
Avrami–Erofeev ;
Prout–Tompkins
Diffusional
1-D
2-D
3-D (Jander)

3-D (Ginstling–Brounshtein)

Contracting geometry
Contracting area ;
Contracting volume ;

34
Table 5. Recent TGA pyrolysis for kinetic studies on various microalgae.

Feedstock TGA operating conditions Kinetic model References


Initial mass Final Heating rate Inert gas flow rate
(mg) temperature (°C min-1) (mL min-1)
(°C)
Dunaliella tertiolecta 10 900 5– 40 50 KFM (Shuping et al., 2010)
Chlorella sp. 100 and 1000 800 15 8500 KFM (Rizzo et al., 2013)
Nannochloropsis
Chlorella vulgaris 10 800 5–40 100 KFM (Agrawal &
Chakraborty, 2013)
Chlorella pyrenoidosa and Spirulina platensis 15 800 10–80 100 KFM (Gai et al., 2013)
Scenedesmus almeriensis, Nannochloropsis 20 1000 40 200 KFM (López-González et al.,
Gaditana and Chlorella vulgaris 2014)
Nannochloropsis oculata 10 1000 5–20 80 KFM (Ceylan & Kazan,
and Tetraselmis sp. 2015)
Chlorella pyrenoidosa 2.5–5 800 5–60 100 KFM (Hu et al., 2015)
Nannochloropsis 4–24 1200 40 50–200 KFM (Sanchez-Silva et al.,
gaditana 2013)
Chlorella protothecoides n/a 800 15–80 60 KFM (Peng et al., 2001)
Chlorococcum humicola n/a 1100 5–20 n/a KFM (Kirtania &
Bhattacharya, 2012b)
Isochrysis sp. and Chlorella sp. 5 900 5–25 100 KFM (Zhao et al., 2015)
Dunaliella tertiolecta 10 800 5–40 50 KFM (Wu et al., 2015)
D. tertiolecta and E. prolifra 10 1000 5–40 100 KFM (Wu et al., 2014)
C. sorokiniana 21 and Monoraphidium 3s35 5 600 20–50 150 KFM (Yang et al., 2014)
Chlorella vulgaris 6 1000 10–40 100 KFM (Chen et al., 2012a)
Saccharina japonica 25 800 10–20 25 KFM (Kim et al., 2012)
Sagarssum sp. 25 800 10–20 25 KFM (Kim et al., 2013)
Consortia grown 5 800 5–40 30 MPRM (Sharara et al., 2014)
Oil extracted Chlamydomonas sp. JSC4 and 5 727 20 100 MPRM (Bui et al., 2016)
Chlorella sorokiniana CY1
Chlorella vulgaris ESP-31 5 700 10 100 MPRM (Bach & Chen, 2017)
KFM: kinetic-free model; MPRM: multiple parallel reaction model

35
Table 6. Components included in one-step kinetic models.

Reaction mechanism Components included


One reaction Whole microalga
Two parallel reactions Carbohydrate & Protein
Lipid
Three parallel reactions Carbohydrate
Protein
Lipid
Four parallel reactions Carbohydrate
Protein
Lipid
Others
Seven parallel reactions Low thermal resistant carbohydrate
High thermal resistant carbohydrate
Low thermal resistant protein
High thermal resistant protein
Lipid
Others
Intermediate products

36
(a)
100
S. obliquus CNW-N
C. sp. JSC4
Chlorella vulgaris ESP-31

80
TGA (%)

60

40

20
200 400 600 800
o
Temperature ( C)
(b)
1

0.8 Carbohydrates
and proteins
DTG (% / C)

0.6
o

0.4
Lipids

0.2

0
200 400 600 800
o
Temperature ( C)

Figure 1. Pyrolytic (a) TGA and (b) DTG curve of three different microalgae.

37
Figure 2. A schematic of microalgae pyrolysis methods and bio-oil and biochar yields.

38
(a) FWO method
4
=0.1
=0.2
=0.3
=0.4
3.5 =0.5
=0.6
=0.7
=0.8
ln

2.5

2
1.4 1.5 1.6 1.7 1.8 1.9
1/T

(b) KAS method


-8.5
=0.1
=0.2
=0.3
=0.4
-9 =0.5
=0.6
=0.7
=0.8
ln(/T )
2

-9.5

-10

-10.5
1.4 1.5 1.6 1.7 1.8 1.9
1/T

Figure 3. Demonstration for data extraction from kinetic-free model (Tran et al., 2014).

39
(a) One reaction (c) Four parallel reactions

2 Exp. 2
Exp.
Conversion rate (d/dt x 10 , s )

Conversion rate (d/dt x 10 , s )


-1

-1
Cal. Cal
Carbohydrate
3

3
Protein
1.5 1.5 Lipid
Others

1 1

0.5 0.5

0 0
100 200 300 400 500 600 100 200 300 400 500 600
o o
Temperature ( C) Temperature ( C)

(b) Two parallel reactions (d) Seven parallel reactions

2 Exp. 2
Conversion rate (d/dt x 10 , s )

Conversion rate (d/dt x 10 , s )

Exp.
-1

-1

Cal. Cal
LTSC Carbohydrate I
3

HTSC Carbohydrate II
1.5 1.5 Protein I
Protein II
Lipid
Others
Intermediates
1 1

0.5 0.5

0 0
100 200 300 400 500 600 100 200 300 400 500 600
o o
Temperature ( C) Temperature ( C)

Figure 4. Improvement in model fit quality with increasing the number of parallel reactions: (a)
one, (b) two, (c) four, and (d) seven reactions (Bach & Chen, 2017).

40
Highlights

1. A state-of-the art review on recent research activities in microalgae pyrolysis is given.

2. Pyrolysis characteristics and kinetics of microalgae via TGA is reviewed.

3. Kinetic-free, single reaction and multiple parallel reaction and distributed activation energy

models are introduced.

4. The kinetic models predicting the thermal degradation of microalgae are examined.

5. Pros and cons of microalgae pyrolysis using TGA are illustrated.

41

You might also like