You are on page 1of 13

Bioresource Technology 246 (2017) 88–100

Contents lists available at ScienceDirect

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

Review

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, Viet Nam
b
Department of Aeronautics and Astronautics, National Cheng Kung University, Tainan 701, Taiwan

h i g h l i g h t s

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


 Pyrolysis characteristics and kinetics of microalgae via TGA is reviewed.
 Kinetic-free, single reaction and multiple parallel reaction and distributed activation energy models are introduced.
 The kinetic models predicting the thermal degradation of microalgae are examined.
 Pros and cons of microalgae pyrolysis using TGA are illustrated.

a r t i c l e i n f o a b s t r a c t

Article history: Pyrolysis is a promising route for biofuels production from microalgae at moderate temperatures (400–
Received 24 May 2017 600 °C) in an inert atmosphere. Depending on the operating conditions, pyrolysis can produce biochar
Received in revised form 15 June 2017 and/or bio-oil. In practice, knowledge for thermal decomposition characteristics and kinetics of microal-
Accepted 16 June 2017
gae during pyrolysis is essential for pyrolyzer design and pyrolysis optimization. Recently, the pyrolysis
Available online 19 June 2017
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
Keywords:
pyrolysis in a kinetic control regime. In addition, a number of kinetic models have been applied to process
Microalgal biomass
Pyrolysis and torrefaction
the TGA data for kinetic evaluation and parameters estimation. This paper aims to provide a state-of-the
Kinetics art review on recent research activities in pyrolysis characteristics and kinetics of various microalgae.
Thermogravimetric analysis Common kinetic models predicting the thermal degradation of microalgae are examined and their pros
Biochar and bio-oil and cons are illustrated.
Ó 2017 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
2. Composition and thermal characteristic of microalgae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
2.1. Microalgae and their chemical composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
2.2. Thermal degradation characteristics of microalgae. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3. Pyrolysis of microalgae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.1. Pyrolysis of microalgae and classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.2. Biochar production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
3.3. Bio-oil production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
4. Thermogravimetric analysis (TGA) technique for pyrolysis kinetics of microalgae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.1. Principle of TGA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.2. Fundamental kinetic expressions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.3. Kinetic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.3.1. Kinetic-free models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

⇑ Corresponding author.
E-mail addresses: weihsinchen@gmail.com, chenwh@mail.ncku.edu.tw (W.-H. Chen).

http://dx.doi.org/10.1016/j.biortech.2017.06.087
0960-8524/Ó 2017 Elsevier Ltd. All rights reserved.
Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100 89

Nomenclature

kðTÞ or k Reaction rate constant (s1)


Abbreviations m Sample mass at any time (g)
TGA Thermogravimetric analysis m0 Initial sample mass (g)
DTG Differential thermogravimetric mf Final residual mass (g)
FWO Flynn-Wall-Ozawa method R Universal gas constant (=8.314 J.mol1.K1)
KAS Kissinger-Akahira-Sunose method T Absolute temperature (K)
DAEM Distributed activation energy model t Conversion time (s)
v Mass of released volatiles at any time (g)
Symbols vf Total mass of released volatiles (g)
A Pre-exponential factor (s1) a Conversion degree
b Heating rate (K/min)
c Contribution factor
Ea Activation energy (kJ/mol) r Deviation of activation energy in DAEM (kJ/mol)
E0 Mean activation energy in DAEM (kJ/mol)

4.3.2. Single reaction model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95


4.3.3. Multiple parallel reaction models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.3.4. Series reaction or consecutive-reaction model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.3.5. Distributed activation energy model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.4. Progress in pyrolysis kinetics study for microalgae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.5. Potential for process design and up-scaling using kinetic data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5. Challenges and opportunities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
6. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

1. Introduction be categorized into first, second, third, and fourth generation biofu-
els. Biofuels produced from food crops, say, sucrose- and starch-
Since the industrial revolution occurred in eighteenth century, a derived bioethanol, belong to first generation biofuels (Chen
considerable amount of coal has been burned for heat and power et al., 2015a). Second generation biofuels are produced from non-
generation as well as industrial applications (Du and Chen, food crops such as lignocellulosic biomass. Biofuels using algal bio-
2006). The consumption of the carbon-based fuels leads to the mass such as microalgae and macroalgae as feedstocks are termed
mass emissions of anthropogenic carbon dioxide into the atmo- third generation biofuels. As for fourth generation biofuels, the
sphere. Nowadays, rapid accumulation of carbon dioxide in the approach deals with the metabolic engineering of algae from oxy-
atmosphere and the accompanied problems such as global warm- genic photosynthetic microorganisms (Lü et al., 2011). Obviously,
ing and climate change have become serious obstacles for environ- third and fourth generation biofuels involve algae-to-biofuel tech-
ment sustainability (Janković and Schultz, 2017). For this reason, nology so that algal biomass is a promising substitute for biofuel
how to efficiently develop renewable energy and sustainable alter- production.
natives to fossil fuels for mitigating the global warming is receiving Algal biomass possesses a number of advantages over lignocel-
a great deal of attention. A number of renewable sources such as luloses, including abundant distribution, fast growth, high produc-
solar, wind, biomass, geothermal, marine, and hydroelectric ener- tion rate, and high photosynthetic or carbon fixing efficiency.
gies can be employed for green energy (Chen et al., 2012b). Solar Moreover, algal biomass is cultivated in water, even using wastew-
energy and wind energy have a substantial growth lately; however, ater as nutrient sources (Chiu et al., 2015), and thus requires no
electricity produced from solar and wind energy is subject to cli- arable land (Bach et al., 2017a). To obtain biofuels from algal bio-
mate and location, and is relatively difficult to be stored. In con- mass through thermochemical conversion, a variety of routes can
trast, biomass grows worldwide, which is not restricted be utilized, depending on the requirements of product phases.
geometrically and by the weather, and can be easily stored and For example, biochar can be obtained from torrefaction and pyrol-
delivered after it is harvested. In addition, carbon stored in biomass ysis, and bio-oil can be produced via pyrolysis and liquefaction
completely comes from carbon dioxide in the atmosphere. Conse- (Chen et al., 2015a). Regarding biochar production, torrefaction
quently, biofuels derived from biomass are considered as carbon- biochar is mainly applied for solid fuels, whereas pyrolysis biochar
neutral fuels. If biochar produced from biomass is used for soil is also utilized for other purposes such as activated carbon, soil
amendment and remediation, it is even termed carbon negative enhancer, fertilizer, etc. (Manyà, 2012).
material (Glaser et al., 2009) in that biochar is mostly not digestible As far as microalgae pyrolysis is concerned, the feedstock is
to microorganisms and a biochar-based soil amendment could thermally decomposed at a moderate-temperature and in an
serve as a permanent carbon-sequestration agent in soils/subsoil oxygen-free environment. The pressure in the reactor is normally
earth layers for thousands of years (Lee et al., 2010). one atmosphere and the temperature is usually between 400 and
Biomass is currently the fourth largest primary energy source in 600 °C (Chen et al., 2015a). Torrefaction is also operated in an
the world, behind coal, petroleum and natural gas (Chen et al., oxygen-free environment, but at lower temperatures, namely,
2012b). According to the types of biomass feedstock, biofuels can 200–300 °C (Chen et al., 2015b). Consequently, torrefaction is
90 Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100

termed mild pyrolysis, and broadly speaking, pertains to pyrolysis to 50 times higher than terrestrial plants (Raheem et al., 2015).
technology. Compared to combustion, gasification, and liquefac- As a consequence, microalgae-based CO2 biological fixation is a
tion, it appears that pyrolysis is a more flexible process for produc- promising means to mitigate anthropogenic CO2 emissions.
ing biofuels in that main products can be solid biochar or liquid Microalgae can be grouped into prokaryotic microalgae
bio-oil, depending on process operating conditions such as heating (cyanobacteria Chloroxybacteria), eukaryotic microalgae (green
rate, reaction temperature, residence time, and feedstock size. algae Chlorophyta), red algae (Rhodophyta), and diatoms (Bacillar-
To understand the fundamental pyrolysis characteristics of iophta) (Sambusiti et al., 2015). Unlike lignocellulosic biomass
microalgae, thermogravimetric analysis (TGA) has been widely which is mainly composed of cellulose (40–60 wt%), hemicellulose
performed. It is known that the reaction rate is highly related to (20–40 wt%), and lignin (10–25 wt%) (Yang et al., 2007), the main
the reactor design and practical pyrolysis operation. Therefore, chemical composition in microalgae consists of carbohydrates (8–
the chemical kinetics is a crucial topic in microalgae pyrolysis 30 wt%), proteins (40–60 wt%), and lipids (5–60 wt%) (Uggetti
and has been extensively studied. Over the past several decades, et al., 2014); other valuable components such as pigments, anti-
a number of kinetic theories from TGA have been conducted to pre- oxidants, fatty acids, and vitamins are also contained. Chemical
dict the thermal decomposition rate of microalgal biomass during composition of microalgae is high variable, and depends mainly
pyrolysis and to aid pyrolyzer design. Nevertheless, an examina- on species, environmental conditions, and cultivation methods
tion of published literature suggests that the review on the chem- (Sambusiti et al., 2015). Carbohydrates, the fermentable sugars,
ical kinetics of microalgae pyrolysis through TGA remains in microalgae are potential feedstocks for bioethanol production
insufficient so far. For this reason, this paper aims to provide a (Sirajunnisa and Surendhiran, 2016). Some microalgal cell walls
comprehensive overview on recent research and development in are composed of cellulose, mannans, xylans, and sulfated glycans.
pyrolysis characteristics and kinetics using microalgae as feed- These polysaccharides can be chemically or enzymatically broken
stocks. A variety of models, including single reaction, multiple par- down into simple sugars and then converted into bioethanol. Cer-
allel reaction, series or consecutive reaction, and distributed tain oleaginous microalgae pertain to lipid-rich species in nature
activation energy models, in predicting the thermal degradation and can be cultivated at optimal conditions to accumulate substan-
behaviors of components in microalgae will be illustrated, and tial quantities of lipids, making high oil yields (5000–
their pros and cons will be discussed. This review is conducive to 100,000 L ha1 year1) (Singh and Olsen, 2011). These lipids in
biochar and bio-oil production, biomass thermal degradation pre- microalgae can be extracted and converted into fatty acid methyl
diction, and experiment and reactor designs. 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
2. Composition and thermal characteristic of microalgae fertilizers (Brennan and Owende, 2010). The elemental and chem-
ical compositions of some common microalgae are tabulated in
2.1. Microalgae and their chemical composition Table 1.

Microalgae are photosynthetic unicellular or simple- 2.2. Thermal degradation characteristics of microalgae
multicellular microorganisms, and are one of the earth’s most
important natural resources (Sambusiti et al., 2015). They account The recognition of the thermal degradation of carbohydrates,
for approximately 50% of global photosynthetic activity (Chiu et al., proteins, and lipids plays a crucial role in investigating the pyroly-
2015). Because some microalgae double in number within hours, tic characteristics of microalgae. To figure out the thermal decom-
they have a short harvesting cycle, say, less than 10 days position behavior of microalgae, the thermogravimetric analysis
(Razeghifard, 2013). In recent years, the cultivation of microalgae (TGA) and derivative thermogravimetric (DTG) curves of three
has received a great deal of attention for pharmaceutical, nutraceu- different microalgal species, consisting of Scenedesmus obliquus
tical, cosmetic, aquiculture, and biorefinery purposes (Zhu, 2015). (S. obliquus) CNW-N (Chen et al., 2014), Chlorella vulgaris
In addition to growth in natural environments, microalgae can also (C. vulgaris) ESP-31 (Bach and Chen, 2017), and Chlamydomonas
be artificially and commercially cultivated in freshwater, marine, sp. (C. sp.) JSC4 (Nakanishi et al., 2014), are shown in Fig. 1 where
and wastewater systems within open ponds (raceways) and closed the temperature is in the range of 105–800 °C (ignoring the drying
photobioreactors (Sambusiti et al., 2015). Certain microalgae can process), the heating rate is 20 °C min1, and nitrogen at a flow
tolerate and adapt to a wide variety of environmental conditions, rate of 100 mL min1 (STP) is used to sweep the samples.
and can be produced all year round. While microalgae are culti- It can be seen that the microalga S. obliquus CNW-N has a smal-
vated, CO2 in flue gas can be used as a carbon source for their ler weight loss within the investigated temperature range when
growth. Moreover, microalgae are able to mitigate CO2 from 10 compared with C. sp. JSC4 and C. vulgaris ESP-31 (Fig. 1a), revealing

Table 1
Elemental 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.


Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100 91

200 °C, water is removed from microalgae. The mass loss in this
(a) stage depends strongly on the moisture content of feedstock. The
100 second stage, from 200 to 600 °C, is the most reactive and accounts
S. obliquus CNW-N for the majority of the mass loss during pyrolysis. In this stage, all
C. sp. JSC4 microalgal components (carbohydrates, proteins, lipids, and other
Chlorella vulgaris ESP-31 minor components) are decomposed to produce chars and release
80 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).
TGA (%)

60 3. Pyrolysis of microalgae

3.1. Pyrolysis of microalgae and classification

40 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
20 pyrolysis can be classified into four modes: (1) slow pyrolysis, (2)
200 400 600 800
o fast pyrolysis, (3) catalytic pyrolysis, and (4) microwave pyrolysis
Temperature ( C) (Chen et al., 2015a) where three different reactors of fixed bed, flu-
(b) idized bed, and auger reactors can be adopted. In general, fixed bed
reactors are used in slow and microwave pyrolysis, and fluidized
1
bed reactors are utilized in fast pyrolysis. In regard to auger reac-
tors, they have also been applied for catalytic pyrolysis. A number
of studies of microalgae pyrolysis are tabulated in Table 2 in which
0.8 Carbohydrates the four types of pyrolysis, reactors, operating conditions, and bio-
and proteins oil and biochar yields are summarized.
In slow pyrolysis, biomass is converted into bio-oil and biochar
DTG (% / C)

at temperatures of 350–600 °C along with a low heating rate


0.6 (10 °C min1) and a long residence time of hot vapor (10–30 s)
o

(Miao et al., 2004). As can be seen in Table 2, the bio-oil and bio-
char yields from the slow pyrolysis of microalgae are in the ranges
0.4 of 18–55 wt% and 10–40 wt%, respectively. It was reported
Lipids (Demirbasß, 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
0.2 was further increased. In contrast, fast pyrolysis is featured by a
high heating rate (3000–36,000 °C min1) 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
0
200 400 600 800 achieve the high heating rate for high bio-oil yield, small biomass
o particles are normally fed into the fluidized bed pyrolyzers. The
Temperature ( C)
bio-oil and biochar yields from the fast pyrolysis of microalgae
Fig. 1. Pyrolytic (a) TGA and (b) DTG curve of three different microalgae. are in the ranges of 18–72 wt% and 22–63 wt%, respectively.
Microalgae pyrolyzed under the aid of catalyst is termed cat-
alytic pyrolysis where Na2CO3, MgO, ZnCl2, and ZSM-5-based zeo-
that the former is relatively thermally resistant in nature. The DTG lites (e.g. H-ZSM-5, Fe-ZSM-5 Cu-ZSM-5 and Ni-ZSM-5) can be
curves of S. obliquus CNW-N, C. vulgaris ESP-31, and C. sp. JSC4 are utilized as the catalysts (Babich et al., 2011; Campanella and
characterized by a single, two, and three peaks, respectively Harold, 2012). The advantages of using catalytic pyrolysis for
(Fig. 1b). Their biggest peaks, corresponding to the main pyrolysis microalgae are that bio-oil with less oxygenic compounds, lower
or devolatilization process, are located at around 320, 290, and acidity, and higher aromatics, calorific value, and bio-oil yield
280 °C, respectively. These peaks triggered are due to the thermal can be obtained (Suali and Sarbatly, 2012), and catalysts used for
decomposition of carbohydrates and proteins (Chen et al., 2014), the pyrolysis can be recycled into the reactor (Babich et al.,
and the thermal degradation peaks of carbohydrates and proteins 2011). Catalytic pyrolysis is usually operated at temperatures of
normally merge together. The peak of S. obliquus CNW-N accompa- 300–600 °C and catalyst-to-biomass mass ratios of 0.2–5 (Chen
nied by a shoulder or those of C. vulgaris ESP-31, and C. sp. JSC4 et al., 2015a). The bio-oil and biochar yields produced from cat-
accompanied by smaller peaks are the consequence of thermal alytic pyrolysis are approximately in the ranges of 20–58 wt%
degradation of lipids. This arises from the fact that the pyrolytic and 16–48 wt% (Babich et al., 2011; Pan et al., 2010). A schematic
temperatures of lipids are higher than those of carbohydrates of the aforementioned microalgae pyrolysis methods and bio-oil
and proteins (López-González et al., 2015). and biochar yields is shown in Fig. 2.
From TGA and DTG curves, overall, the microalgae pyrolysis can Dielectric materials are able to convert microwaves into ther-
be divided into three different reactive stages, which are separated mal energy through a dielectric heating process (Chen and Lin,
by vertical lines shown in Fig. 1b. In the first stage, which lasts to 2010). Accordingly, bio-oil and biochar can be produced from
92 Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100

microalgae through microwave pyrolysis along with microwave

(Campanella and Harold, 2012)


absorbers (e.g. SiC, Fe3O4, CuO, water, fats, activated carbon, bio-
char, ionic liquids, sulfuric acid, etc). Microwave-assisted heating

(Chiaramonti et al., 2017)


(Francavilla et al., 2015)
possesses a number of merits over traditional heating such as rapid
(Grierson et al., 2009)
(Grierson et al., 2009)
(Grierson et al., 2009)

(Miao and Wu, 2004)


(Jena and Das, 2011)
heating, uniform internal heating of feedstock, instantaneous

(Borges et al., 2014)


(Borges et al., 2014)
(Wang et al., 2017)
(Wang et al., 2015)

(Dong et al., 2015)


(Miao et al., 2004)
(Cho et al., 2015)

(Pan et al., 2010)


response for rapid start-up and shut down, no need for agitation

(Xie et al., 2015)


(Hu et al., 2016)

(Hu et al., 2014)


(Hu et al., 2014)
via fluidization, and hence fewer particles (ashes) in the produced
bio-oil (Borges et al., 2014; Chen et al., 2015a). Microwave pyroly-
Reference

sis is usually operated at temperatures of 450–800 °C and absorber


contents of 5–30 wt%. The bio-oil and biochar yields from micro-
wave pyrolysis are in the ranges of 21–59 wt% and 14–49 wt%,
Biochar

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


34–63
22–36

14–37
26–49
26–46

25–48
28–40

17.77
30.26

Among the pyrolysis products, biochar production from slow


11

16

16
Yield (wt%)

pyrolysis is a mature technology; while bio-oil production are still


Bio-oil

23–29

58–72
24–43
39–43

41–59
41–57
21–31

38–58
55.21

37.11
20.21 being developed, employing several advanced technologies such as
36.8

28.6
33
41
41

18

47
fast, catalytic, and microwave pyrolysis as introduced earlier. Also,
practical applications of biochar are wider than those of bio-oil,
Duration (min)

which will be discussed in the next subsections.

3.2. Biochar production


8.33
8.33
120
15
20
20
20
60

30
30

Biochar produced from biomass can be applied in agriculture


Sweep gas flow rate (mL min1)

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 energy contents of microalgae feedstock, respec-
tively (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
1333
6667
6667

1333
1333
250

180

250
100
100
100

600

400

account of rich carbon contained in biochar which can remain


30

stable in soil for hundred or even thousand years (Chaiwong


Heating rate (°C min1)

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 cap-
ture and storage solution through biomass pyrolysis to produce
biochar or long-term carbon sink, using soil as the storage media
Operating conditions

36,000
36,000
3.5–7

(Grierson et al., 2011). This substantially delays the release of CO2 -


3000
6.67
10
10
10

10

10

into the atmosphere and thereby mitigates global warming. Bio-


char contains most of the feedstock’s mineral components so
Temp. (°C)

450–550

450–550
450–550
350–500

450–600
600–800

500–700

300–500

that it can be used as a soil amendment, and the high inorganic


contents (e.g. potassium, phosphorous, and nitrogen) in biochar
475
530

550
500
500
500

500

500
500

800
800
600

may be suitable to provide nutrients for crop production (Wang


et al., 2013). Moreover, biochar has a highly porous structure.
Fluidized bed
Fluidized bed
Fluidized bed

The addition of bio-char into soil could improve water retention


Fixed bed
Fixed bed
Fixed bed
Fixed bed
Fixed bed
Fixed bed
Fixed bed
Fixed bed

Fixed bed
Fixed bed
Fixed bed
Fixed bed
Fixed bed
Fixed bed
Fixed bed

and increase the surface area of the soil. This facilitates the effi-
Reactor

Auger
Auger

ciency of nutrient use (Chaiwong et al., 2013).

3.3. Bio-oil production


Catalytic (biochar based catalyst)

Bio-oil produced from microalgae pyrolysis has several environ-


mental advantages over fossil oil. For example, bio-oil is CO2 or
greenhouse gas (GHG) neutral so that it generates carbon dioxide
Catalytic(H-ZSM-5)
Catalytic (HZSM-5)

Catalytic (ZnCl2)

credits. Microalgae contain extremely low amount of sulfur (Kim


Catalytic (MgO)
Pyrolysis Mode
A list of literature of microalgae pyrolysis.

et al., 2014); hence there are almost no SOx emissions when


Microwave
Microwave
Microwave

microalgae-based bio-oil is burned. Meanwhile, it was addressed


that pyrolytic bio-oil from microalgae has lower oxygen and water
Slow
Slow
Slow
Slow
Slow
Slow
Slow
Slow
Fast
Fast
Fast
Fast

contents and higher calorific value and carbon content than that
from lignocellulosic biomass (Rizzo et al., 2013). However, it is
Chaetocerous muelleri

Chlorella sorokiniana
Dunaliella tertiolecta

notable that bio-oil from microalgae showed higher nitrogen con-


Spirulina platensis
Chlorella vulgaris

Chlorella vulgaris

Chlorella vulgaris

Chlorella vulgaris
Chlorella vulgaris
C. protothecoides
C. protothecoides

tent compared with petroleum oils, likely due to nitrogen con-


Nannochloropsis

Nannochloropsis
Nannochloropsis

Nannochloropsis
N-heterocyclic

tained in proteins inside the microalgae. This leads to higher NOx


Chlorella like

Chlorella sp.

Green algae
Feedstock

emissions in fuel use and catalyst poisoning in deoxygenation pro-


Chlorella

cess for further processing (Wildschut et al., 2010). Due to the inte-
Table 2

ferior properties compared with petroleum oil, the applications of


pyrolysis bio-oil is still limited now. Nevertheless, bio-oil can be
Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100 93

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

used not only as fuel but also as raw material for chemical produc- under air or inert environment for subsequent kinetic studies
tion. According to Xiu and Shahbazi (2012), industrial applications (Bach et al., 2017b; Broström et al., 2012). The patterns reveal
of bio-oil include: the response of heating rate, the inherent biomass properties and
severity that define the char combustion profile in the form of ther-
 Combustion fuel for heat and power generation. mograph (Islam et al., 2016).
 Transportation fuel after upgrading.
 Production of pharmaceuticals, surfactants, and biodegradable 4.2. Fundamental kinetic expressions
polymers.
 Liquid smoke and wood flavors. It is known that the knowledge of thermal behavior and pyrol-
 Production of adhesives, chemicals, and resins. ysis kinetics of microalgae is very crucial for recognizing the ther-
mal degradation (or stability) and conversion of substance and
4. Thermogravimetric analysis (TGA) technique for pyrolysis product formation; it is also conducive to process rate prediction,
kinetics of microalgae experimental design, operational parameter decision, and practical
operation. In general, the pyrolysis kinetics is expressed in terms of
4.1. Principle of TGA the Arrhenius law which comprises activation energy, frequency
factor, and reaction order. The reaction rate under an isothermal
TGA pertains to a thermal analysis technique in nature in which condition as a function of time is described as:
 
the weight of a sample is continuously measured under a con- da Ea
trolled temperature program. Based on the variation of sample ¼ kðTÞ  f ðaÞ ¼ Aexp  f ðaÞ ð1Þ
dt RT
weight with respect to temperature or time, the differential ther-
mogravimetric (DTG) curve can be further obtained by differenti- where a is the conversion degree, t is the conversion time, A is the
ate the TGA curve where the physical and chemical properties of pre-exponential factor, Ea is the activation energy of the reaction, R
samples as a function of temperature can be determined is the universal gas constant, and T is the absolute temperature. The
(Yahiaoui et al., 2015). The utilization of thermogravimetric ana- conversion degree (a) is defined as the mass fraction of decomposed
lyzer has the advantages of easy operation, minimal quantity of solid or released volatiles:
m0  m v
feedstock, precise control and record of temperature and sample a¼ ¼ ð2Þ
weight loss. TGA has been used to perform the proximate analysis m0  mf v f
of biomass (Saldarriaga et al., 2015) and figure out the thermal where m0 and mf are the initial and final masses of solid, m is the
behavior of microalgae such as combustion (Tang et al., 2011), igni- mass of solid at any time; v f is the total mass of released volatiles,
tion and burnout (Chen et al., 2016), torrefaction (Chen et al., and v is the mass of released volatiles at any time.
2014), pyrolysis (Sanchez-Silva et al., 2013), gasification The reaction rate in Eq. (1) can be easily transformed into a non-
(Sanchez-Silva et al., 2013), and chemical kinetics (Bach and isothermal expression, which describes the conversion rate as a
Chen, 2017). The relevant studies using TGA as a tool are shown function of temperature at a constant heating rate, b, as the
in Table 3. following:
TGA has received immense attention in understanding solid fuel
degradation to release energy. TGA is able to trace the thermal da dt da 1 da
¼ ¼ ð3Þ
degradation patterns of fresh biomass, biochar, and hydrochar dT dT dt b dt
94 Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100

By substituting Eq. (1) into Eq. (3), the non-isothermal rate

(Kirtania and Bhattacharya, 2012a)


expression is obtained as:
 
da kðTÞ A Ea

(López-González et al., 2015)


¼  f ðaÞ ¼ exp  f ðaÞ ð4Þ

(Sanchez-Silva et al., 2013)


dT b b RT
(Kebelmann et al., 2013)

(Francavilla et al., 2015)


(Bach and Chen, 2017)
(Marcilla et al., 2009)
The function f ðaÞ in Eqs. (1) and (4) can be expressed by several

(Rizzo et al., 2013)


(Chen et al., 2016)

(Chen et al., 2014)


(Tang et al., 2011)

(Wu et al., 2014)


equations depending on the selection of reaction mechanisms. A

(Li et al., 2017)


list of expressions for f ðaÞ is adopted from (White et al., 2011)
Reference

and presented in Table 4. Among those, mechanisms based on


reaction order are commonly used for most biomass pyrolysis.

4.3. Kinetic models


Duration (min)

This part reviews common pyrolysis kinetic models that have


been successfully applied for microalgae. Like other biomass mate-
rials, microalgae pyrolysis virtually consists of numerous chemical
60
60

60

reactions, and a number of intermediates and substances are pro-


duced during the process. In the simplest kinetics approach, pyrol-
550, 650, 750, 850

ysis activation energy and pre-exponential factor can be


Temperature (°C)

approximately calculated from TGA data without any knowledge


about the reaction mechanism. These approximations are called
125–1000
100–1000

25–1100
105–800

105–800

105–700
25–900

50–750
25–900

25–800

25–800
25–100

‘‘kinetic-free model” because they offer very limited kinetic infor-


mation, except the activation energy and pre-exponential factor.
On the other hand, more comprehensive kinetic approaches pro-
pose that microalgae produce chars and volatiles during pyrolysis,
Heating rate (°C min1)

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 prod-
ucts in only one stage. In another pyrolysis kinetic model, it
0.5, 1.0, 1.5

assumes that microalgae form intermediates and initial volatiles


in the primary reaction. Then, the intermediates are converted into
5–40
5–20
100

35

15
20

20

20

10
10

10

40
10

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
Pyrolysis under different CO2 concentrations with TG-IR analysis

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
Combustion, ignition and burnout (in air)

energy and pre-exponential factor from experimental TGA data at


any conversion rates through linear transformations of Eq. (4). By
Pyrolysis characteristics and kinetics

Pyrolysis characteristics and kinetics

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


 
Pyrolysis with TG-FTIR analysis

da A Ea
Non-isothermal torrefaction

¼ exp  dT ð5Þ
f ðaÞ b RT
Isothermal torrefaction

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


Thermal behavior

Pyrolysis kinetics

Pyrolysis kinetics

Z  Z T  
Pyrolysis (in N2)
Investigation on thermal behaviors of microalgae using TGA.

T
da A Ea
gðaÞ ¼ ¼ exp dT ð6Þ
Combustion

Combustion

Gasification

T0 f ðaÞ b T0 RT
Pyrolysis

Pyrolysis

Pyrolysis

Pyrolysis

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.
Dunaliella tertiolecta and its residue
Chlamydomonas sp. JSC4 residue

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


     
b AR Ea
ln ¼ ln  ð7Þ
Nannochloropsis gaditana

T2 Ea gðaÞ RT
Chlorella protothecoides

Chlorococcum humicola
Scenedesmus obliquus
Chlorella vulgaris

Nannochloropsis sp.

   
Nannochloropsis
Tetraselmis suecica

Chlorella vulgaris

AEa Ea
Scenedesmus sp.

lnb ¼ ln  1:0516 ð8Þ


RgðaÞ
Chlorella spp.

RT
C. reinhardtii

D. tertiolecta
Microalgae

At any conversion rate (a), the plot of lnðTb2 Þ or lnb versus


Table 3

inverted temperature ð1T Þ should create straight lines, from which


the activation energy (Ea ) and the pre-exponential factor (A) can
Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100 95

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

Reaction mechanism f ðaÞ


Reaction order
Zero order 1
First order ð1  aÞ
nth order ð1  aÞn
Nucleation
Power law nðaÞð11=nÞ
Exponential law ln a
ð11=nÞ
Avrami–Erofeev nð1  aÞ½ ln að1  aÞ ; (n=1,2,3,4)
Prout–Tompkins að1  aÞ
Diffusional
1-D 1=2a
1
2-D ½lnað1  aÞ
3-D (Jander) 1
2 ð1
3
 aÞ 2=3
½1  ð1  aÞ1=3 
3-D (Ginstling–Brounshtein) 1
2 ½ð1
3
 aÞ1=3  1
Contracting geometry
Contracting area ð1  aÞð11=nÞ ; n ¼ 2
Contracting volume ð1  aÞð11=nÞ ; n ¼ 3

be calculated through the slope and the intercept, respectively.


Generally, the conversion rates vary from 0.1 to 0.9 with an inter-
val of 0.1. Plotting of these equations is presented in Fig. 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.
k
Microalga ! Char þ Volatile ð9Þ
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 based on a three parallel reaction
model, which has been successfully applied for lignocellulosic bio-
mass for decades (Di Blasi, 2008; Orfão et al., 1999). Due to differ-
ence 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 microalgae pyrolysis
(Bach and 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.
8 Fig. 3. Demonstration for data extraction from kinetic-free model (Tran et al.,
>
>
k1
Comp 1 ! Char 1 þ Volatile 1
2014).
>
>
>
> A microalga generally consists of three main components (car-
< Comp 2 k!
2
Char 2 þ Volatile 2
Microalga ð10Þ bohydrate, protein, and lipid) and other minor components. Due
>
> ..
>
> . to different thermal reactivities, these components behave differ-
>
>
: kN ently during pyrolysis. Thus, appropriate models should include
Comp N ! Char N þ Volatile N three (excluding other components) or four (including other com-
If the microalgal components are divided into two groups based ponents) parallel reactions to better describe the pyrolysis of
on their thermal stability, a two-reaction model can be expressed microalga, as shown in Eq. (12).
as: 8
8 >
>
k1
Carbohydrate ! Char 1 þ Volatile 1
< >
>
k1
LTSC ! Char 1 þ Volatile 1 >
>
Microalga ð11Þ < k2
: Protein ! Char 2 þ Volatile 2
k2
HTSC ! Char 2 þ Volatile 2 Microalga ð12Þ
>
> k
Lipid ! Char 3 þ Volatile 3
3
>
>
>
>
where LTSC and HTSC designate low thermal stable components : k4
Others ! Char 4 þ Volatile 4
and high thermal stable components in microalgae, respectively
(Chen et al., 2015a). Considering the thermal stability of microalgal Furthermore, a more detailed model may contain up to seven
components, the former presents carbohydrates and proteins, while parallel reactions. In this model, carbohydrate and protein in a
the latter stands for lipids (Bach and Chen, 2017). microalga are partitioned into two groups: high and low thermal
96 Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100

resistant substances. Another new component in this complex a distributed activation energy model (DAEM) can employed for
model is the intermediate products, which are formed during the pyrolysis kinetic study. In the DAEM, complex pyrolysis decompo-
pyrolysis of a microalga at low temperatures and react indepen- sition of any microalagal component can be described by a series of
dently with other components. reactions having different activation energy values but the same
pre-exponential factor. A general equation for the DAEM can be
4.3.4. Series reaction or consecutive-reaction model written as:
Different from the aforementioned parallel reaction models, Z  Z t 
v 1
Ea
this model consists of reactions in series: initial microalga is con- 1 ¼ exp A e RT dt f ðEÞdE ð15Þ
verted to Intermediate and Volatile 1 in Eq. (13); thereafter, the
vf 0 0

Intermediate forms final Char and Volatile 2 in Eq. (14). This series The f ðEÞ is the distribution function of activation energy, which
reaction model was most applied for the isothermal kinetic study can be described by Gaussian, Weibull, or Gamma distribution (de
(Bach et al., 2016; Bates and Ghoniem, 2012). Caprariis et al., 2012). Among those, the Gaussian function (Eq
k1 (16)), which show a mean value (E0) and its deviation r, is used
Microalga ! Intermediate þ Volatile 1 ð13Þ more commonly.
!
k2 1 ðE  E0 Þ2
Intermediate ! Char þ Volatile 2 ð14Þ f ðEÞ ¼ pffiffiffiffiffiffiffi exp  ð16Þ
r 2p 2r2
It should be kept in mind that the conversion of the intermedi-
ate in the secondary reaction depends on the degradation of initial Although the DAEM offers much better fit quality than other
microalga in the primary reaction. This assumption is different models, the model has a double-layer integral and one variable
from the aforementioned seven parallel reaction model, where (E) goes from zero to infinite (Eq. (15)), which cannot be calculated
intermediate products react independently with other compo- directly. However, some simplifications (Miura, 1995; Please et al.,
nents. (Branca and Di Blasi, 2004) compared the parallel and series 2003) were proposed to reduce the complexity of the model in
reaction models for lignocellulose biomass and found that kinetic practical calculation.
parameters obtained in consecutive combustion were invariant
with the selection of model. However, no similar work for microal- 4.4. Progress in pyrolysis kinetics study for microalgae
gae has been reported.
Currently, the number of kinetic studies on microalgae pyroly-
4.3.5. Distributed activation energy model sis via TGA is much smaller than that of pyrolysis studies for
All the aforementioned models assume that the activation bio-oil and biochar production in other reactors (e.g. fixed bed,
energy values of the microalgal components are constant but a fluidized bed, auger, etc.). An up-to-date list of publications for
component may consist of several reacting species whose reactiv- TGA pyrolysis studies are presented in Table 5. From the table, typ-
ities are different. These variations can lead to the fact that the ical ranges for TGA operating conditions can be established. TGA
activation energy of a component is not a constant value but in a pyrolysis normally requires amounts of 5–20 mg microalga and
range of values (Várhegyi et al., 2010). On accounting for this issue, inert gas (nitrogen or argon) flow rates of 50–200 mL min1.

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

Feedstock TGA operating conditions Kinetic References


model
Initial mass Final Heating Inert gas flow
(mg) temperature rate rate
(°C) (°C min1) (mL min1)
Dunaliella tertiolecta 10 900 5– 40 50 KFM (Shuping et al., 2010)
Chlorella sp. 100 and 800 15 8500 KFM (Rizzo et al., 2013)
Nannochloropsis 1000
Chlorella vulgaris 10 800 5–40 100 KFM (Agrawal and Chakraborty,
2013)
Chlorella pyrenoidosa and Spirulina platensis 15 800 10–80 100 KFM (Gai et al., 2013)
Scenedesmus almeriensis, Nannochloropsis Gaditana and 20 1000 40 200 KFM (López-González et al., 2014)
Chlorella vulgaris
Nannochloropsis oculata and Tetraselmis sp. 10 1000 5–20 80 KFM (Ceylan and Kazan, 2015)
Chlorella pyrenoidosa 2.5–5 800 5–60 100 KFM (Hu et al., 2015)
Nannochloropsis gaditana 4–24 1200 40 50–200 KFM (Sanchez-Silva et al., 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 and 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 Chlorella 5 727 20 100 MPRM (Bui et al., 2016)
sorokiniana CY1
Chlorella vulgaris ESP-31 5 700 10 100 MPRM (Bach and Chen, 2017)

KFM: kinetic-free model; MPRM: multiple parallel reaction model


Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100 97

Moreover, heating rates of 5–40 °C/min can be applied to reach Table 6


final temperatures at 700–1000 °C. Because all works in Table 5 Components included in different one-step kinetic models.

employed non-isothermal mode for TGA pyrolysis in order to Reaction mechanism Components included
ensure completed pyrolysis, the final temperatures in these studies One reaction Whole microalga
are virtually higher than those for pyrolysis in other reactors Two parallel reactions Carbohydrate & Protein
(Table 2), which were carried out under an isothermal mode at Lipid
the final temperatures or at very high heating rates. Three parallel reactions Carbohydrate
Protein
Although various methods has been applied to study pyrolysis Lipid
kinetics of microalgae, most of the works listed in Table 5 Four parallel reactions Carbohydrate
employed kinetic-free models, presented in Section 4.3.1, to calcu- Protein
late a range of pyrolysis activation energy and pre-exponential fac- Lipid
Others
tor at various conversion rates (normally, from 0.1 to 0.9 with
Seven parallel reactions Low thermal resistant carbohydrate
intervals of 0.1). (Shuping et al., 2010) reported that pyrolysis acti- High thermal resistant carbohydrate
vation energy of microalga Dunaliella tertiolecta did not show large Low thermal resistant protein
variations when calculated by Kissinger–Akahira–Sunose (KAS) High thermal resistant protein
and Flynn–Wall–Ozawa (FWO) methods. The activation energy Lipid
Others
values were 131.7–152.7 kJ/mol by KAS method and 134.4–
Intermediate products
152.9 kJ/mol by FWO method, respectively. Similarly, (Agrawal
and 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 isothermal mode (Bach et al., 2016; Bates and Ghoniem, 2012),
method, researchers revealed that the activation energy of Chlorella rather than common non-isothermal TGA employed in most recent
spp. were 71.3–79.2 kJ/mol (Rizzo et al., 2013), and those of S. pyrolysis studies. Although it can be found that (Sanchez-Silva
Platensis and C. Potothecoides were 76–97 kJ/mol and 42–52 kJ/mol et al., 2013) employed series reactions to investigate the release
(Peng et al., 2001). In addition, (Gai et al., 2013) employed of gases during pyrolysis, it requires additional mass spectrometry
Vyazovkin method and showed that activation energy of Chlorella to record the evolution of produced gases and not really employs
pyrenoidosa and Spirulina platensis were 8.9–114.5 kJ/mol and TGA technique.
74.4–140.1 kJ/mol, respectively. Due to limited investigation Compared to multiple parallel reaction models, application of
points, these models are unable to reproduce thermogravimetric the DAEM for microalgal pyrolysis is less common. Although the
curves or provide any information about the fit quality between DAEM is widely applied for kinetic studies of several lignocellulosic
the modeled and experimental TGA data for model evaluation. biomass species, only a few works (Ceylan and Kazan, 2015; Hu
Few studies divided DTG curves into different zones (Agrawal et al., 2015) employed simplified DAEM for pyrolysis kinetics of
and Chakraborty, 2013) or decreased conversion rate intervals microalgae N. oculata, Tetraselmis sp., Chlorella pyrenoidosa and
(down to 0.01) (Hu et al., 2015) to increase the number of investi- bloom-forming cyanobacteria. However, very little information
gation points, which can approximately draw predicted curves. In about the advantages of the DAEM can be obtained from these
other study, (Sanchez-Silva et al., 2013) employed regression studies despite that the authors concluded that the DAEM was bet-
analysis to interpolate additional points and provided statistical ter than a single-step model and it gave excellent fits between sim-
R2 values to evaluate their model. Though some improvements ulated results and experimental data. In addition, Hu et al. (2015)
and modifications have been applied for the kinetic-free models found that activation energy values estimated from the DAEM
to help create modeled curves and provide fit quality information, were always lower than those from the single-step global model.
these models cannot provide the reaction mechanism, e.g. how In details, the activation energy of Chlorella pyrenoidosa and
microalgal components behave during pyrolysis. To overcome this bloom-forming cyanobacteria from the DAEM were respectively
issue, multiple parallel reaction models (Section 4.3.3) are more 100.6 and 144.5 kJ/mol, while these values from the single-step
appropriate. The number of independent reactions can be four global model were 143.71 and 173.46 kJ/mol.
(Sharara et al., 2014), five (Bui et al., 2016), or up to seven (Bach
and Chen, 2017). Moreover, each reaction represents the degrada- 4.5. Potential for process design and up-scaling using kinetic data
tion of each microalgal component. (Sharara et al., 2014) assumed
that component 1 includes moisture and light hydrocarbons, com- Seeing that pyrolysis is an important process for producing bio-
ponent 2 and 3 are respectively protein and starch, while compo- oil and biochar from microalgae, it has received growing attention
nent 4 represent passive pyrolysis weight loss. On the other recently. Pyrolysis kinetics study thus plays an important role in
hand, (Bui et al., 2016) assigned five components as hemicellulose, the design and optimization of pyrolyzers and it is worth to inves-
cellulose, lignin, lipid, and protein. Recently, (Bach and Chen, 2017) tigate the kinetics prior to practical applications (Doyle, 1961).
examined different one-step models and changed the number of According to (Hakvoort et al., 1989), the thermal conversion curves
involved reactions, from one to seven. The identical names of the may differ for identical experiment conditions but the average acti-
components in these models are summarized in Table 6. They com- vation energy should be nearly constant despite different heating
pared the fit quality of the models with increasing the number of rates and heating programs. Therefore, TGA becomes the most
reactions and found that the more reactions the model has, the favorable tool for kinetics studies of biomass materials including
better fit quality it offers. Improvement in fit quality with increas- microalgae for several decades (Várhegyi, 2007), even though heat-
ing the number of reactions is visually demonstrated in Fig. 4. It ing profile in TGA experiment is far from that for practical
can be seen that the unique advantage of the multiple reaction situations.
models compared with the kinetic-free models is that the former Currently, pyrolysis of microalgae has not been exploited at
can provide the information about the reactivity of each industrial scale yet. However, kinetic data from the available stud-
component in a microalga. ies could be beneficial for the process design and up-scaling. Prior
Recently, the series reaction model (Section 4.3.4) has not been to industrial deployment, it needs to study conceptual process
applied for microalga kinetic study yet. It may be due to the modeling and optimization using commercial simulators (e.g.,
fact that this model is more suitable to investigate TGA data in Aspen Plus). Modeling of pyrolysis can employ kinetic reaction
98 Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100

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

models based on thermodynamic equilibrium calculations, which studies, a study (Bach and Chen, 2017) provided comprehensive
can offer flexible and predictive simulations for a wide range of kinetic information but only limited to Chlorella vulgaris ESP-31,
biomass feedstock (Peters et al., 2017). Moreover, these kinetic applications of the models for other species are strongly recom-
schemes use power law kinetic expressions, which is quite similar mended. Moreover, some conflictions are found among the studies
to the fundamental Arrhenius’ equation and has been widely used when assigning reactions to components. Addressing this issue in
in pyrolysis kinetic studies. Results obtained from these models further works is also advisable. Last but not least, application of
can provide an overview on the pyrolysis processes at plant level the DAEM for microalgae pyrolysis kinetic study needs more atten-
as well as assessments on energy, economic and even environmen- tion due to the excellent fit of the DAEM compared to other
tal impacts of the process. In short, kinetic study is always the first models.
step and essential to provide information for further process design However, it should be reminded that this technique aims at
and up-scaling. 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 temper-
5. Challenges and opportunities atures must be employed to establish a chemical kinetic control
regime. Depending on feedstock characteristics, different sample
Due to the limited number of publications reporting on multiple weights can be found in the literature. However, an amount of
parallel reaction models, it suggests that further kinetic studies 5 mg or less with particles sizes less than 150 mm are recom-
would employ these models due to their advantages. Among the mended for TGA pyrolysis study. In addition, a final temperature
Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100 99

in the range of 700–800 °C can be employed for non-isothermal Chen, W.-H., Lin, B.-J., 2010. Effect of microwave double absorption on hydrogen
generation from methanol steam reforming. Int. J. Hydrogen Energy 35 (5),
mode. In practical applications, complex interactions between
1987–1997.
the chemistry and transport phenomena always occur at inter- Chen, C., Ma, X., He, Y., 2012a. Co-pyrolysis characteristics of microalgae Chlorella
and intra-particles. Thus it recommends exploring more detailed vulgaris and coal through TGA. Bioresour. Technol. 117, 264–273.
models coupling with heat, mass and momentum transports to fur- Chen, W.-H., Lu, K.-M., Tsai, C.-M., 2012b. An experimental analysis on property and
structure variations of agricultural wastes undergoing torrefaction. Appl.
ther understand the behaviors of microalgal particles during prac- Energy 100, 318–325.
tical pyrolysis. Chen, W.-H., Wu, Z.-Y., Chang, J.-S., 2014. Isothermal and non-isothermal
torrefaction characteristics and kinetics of microalga Scenedesmus obliquus
CNW-N. Bioresour. Technol. 155, 245–251.
Chen, W.-H., Lin, B.-J., Huang, M.-Y., Chang, J.-S., 2015a. Thermochemical conversion
6. Conclusions of microalgal biomass into biofuels: a review. Bioresour. Technol. 184, 314–327.
Chen, W.-H., Peng, J., Bi, X.T., 2015b. A state-of-the-art review of biomass
A comprehensive overview on recent research activities in TGA torrefaction, densification and applications. Renew. Sustain. Energy Rev. 44,
847–866.
pyrolysis characteristics and kinetics of various microalgae has Chen, Y.-C., Chen, W.-H., Lin, B.-J., Chang, J.-S., Ong, H.C., 2016. Impact of torrefaction
been presented, which is conducive to pyrolyzer design, operation on the composition, structure and reactivity of a microalga residue. Appl.
optimization, and biofuel production. For kinetics, the kinetic-free Energy 181, 110–119.
Chiaramonti, D., Prussi, M., Buffi, M., Rizzo, A.M., Pari, L., 2017. Review and
models are simple, but provide very limited kinetic information.
experimental study on pyrolysis and hydrothermal liquefaction of microalgae
Therefore, multiple parallel reaction models are more preferable for biofuel production. Appl. Energy 185 (Part 2), 963–972.
for further intrinsic kinetic studies, which are able to clarify the Chiu, S.-Y., Kao, C.-Y., Chen, T.-Y., Chang, Y.-B., Kuo, C.-M., Lin, C.-S., 2015.
reactions of individual microalgal components. Last but not least, Cultivation of microalgal Chlorella for biomass and lipid production using
wastewater as nutrient resource. Bioresour. Technol. 184, 179–189.
more attention on the DAEM and detailed models coupling with Cho, S.-H., Kim, K.-H., Jeon, Y.J., Kwon, E.E., 2015. Pyrolysis of microalgal biomass in
heat, mass and momentum transports to further understand the carbon dioxide environment. Bioresour. Technol. 193, 185–191.
pyrolysis behaviors of microalgal particles in practical situations de Caprariis, B., De Filippis, P., Herce, C., Verdone, N., 2012. Double-gaussian
distributed activation energy model for coal devolatilization. Energy Fuels 26
are recommended. (10), 6153–6159.
Demirbasß, A., 2006. Oily products from mosses and algae via pyrolysis.
Energy Sources Part A: Recovery Utilization Environmental Effects 28 (10),
Acknowledgements 933–940.
Di Blasi, C., 2008. Modeling chemical and physical processes of wood and biomass
pyrolysis. Prog. Energy Combust. Sci. 34 (1), 47–90.
The authors acknowledge the financial support from the Min- Dong, T., Gao, D., Miao, C., Yu, X., Degan, C., Garcia-Pérez, M., Rasco, B., Sablani, S.S.,
istry of Science and Technology, Taiwan, R.O.C., under the contract Chen, S., 2015. Two-step microalgal biodiesel production using acidic catalyst
MOST 106-2923-E-006-002-MY3 for this research. The authors generated from pyrolysis-derived bio-char. Energy Convers. Manage. 105,
1389–1396.
also thank Prof. Jo-Shu Chang at National Cheng Kung University Doyle, C.D., 1961. Kinetic analysis of thermogravimetric data. J. Appl. Polym. Sci. 5
for providing the microalgae in this study. (15), 285–292.
Du, S.-W., Chen, W.-H., 2006. Numerical prediction and practical improvement of
pulverized coal combustion in blast furnace. Int. Commun. Heat Mass Transfer
References 33 (3), 327–334.
Francavilla, M., Kamaterou, P., Intini, S., Monteleone, M., Zabaniotou, A., 2015.
Cascading microalgae biorefinery: fast pyrolysis of Dunaliella tertiolecta lipid
Agrawal, A., Chakraborty, S., 2013. A kinetic study of pyrolysis and combustion of
extracted-residue. Algal Res. 11, 184–193.
microalgae Chlorella vulgaris using thermo-gravimetric analysis. Bioresour.
Gai, C., Zhang, Y., Chen, W.-T., Zhang, P., Dong, Y., 2013. Thermogravimetric and
Technol. 128, 72–80.
kinetic analysis of thermal decomposition characteristics of low-lipid
Babich, I.V., van der Hulst, M., Lefferts, L., Moulijn, J.A., O’Connor, P., Seshan, K., 2011.
microalgae. Bioresour. Technol. 150, 139–148.
Catalytic pyrolysis of microalgae to high-quality liquid bio-fuels. Biomass
Glaser, B., Parr, M., Braun, C., Kopolo, G., 2009. Biochar is carbon negative. Nature
Bioenergy 35 (7), 3199–3207.
Geosci. 2 (1). 2-2.
Bach, Q.-V., Chen, W.-H., 2017. A comprehensive study on pyrolysis kinetics of
Grierson, S., Strezov, V., Ellem, G., McGregor, R., Herbertson, J., 2009. Thermal
microalgal biomass. Energy Convers. Manage. 131, 109–116.
characterisation of microalgae under slow pyrolysis conditions. J. Anal. Appl.
Bach, Q.-V., Chen, W.-H., Chu, Y.-S., Skreiberg, Ø., 2016. Predictions of biochar yield
Pyrol. 85 (1–2), 118–123.
and elemental composition during torrefaction of forest residues. Bioresour.
Grierson, S., Strezov, V., Shah, P., 2011. Properties of oil and char derived from slow
Technol. 215, 239–246.
pyrolysis of Tetraselmis chui. Bioresour. Technol. 102 (17), 8232–8240.
Bach, Q.-V., Chen, W.-H., Lin, S.-C., Sheen, H.-K., Chang, J.-S., 2017a. Wet torrefaction
Hakvoort, G., Schouten, J.C., Valkenburg, P.J.M., 1989. The determination of coal
of microalga Chlorella vulgaris ESP-31 with microwave-assisted heating. Energy
combustion kinetics with thermogravimetry. J. Therm. Anal. Calorim. 35 (2),
Convers. Manage. 141, 163–170.
335–346.
Bach, Q.-V., Tran, K.-Q., Skreiberg, Ø., 2017b. Comparative study on the thermal
Hu, Z., Ma, X., Li, L., Wu, J., 2014. The catalytic pyrolysis of microalgae to produce
degradation of dry- and wet-torrefied woods. Appl. Energy 185 (Part 2), 1051–
syngas. Energy Convers. Manage. 85, 545–550.
1058.
Hu, M., Chen, Z., Guo, D., Liu, C., Xiao, B., Hu, Z., Liu, S., 2015. Thermogravimetric
Bates, R.B., Ghoniem, A.F., 2012. Biomass torrefaction: modeling of volatile and solid
study on pyrolysis kinetics of Chlorella pyrenoidosa and bloom-forming
product evolution kinetics. Bioresour. Technol. 124, 460–469.
cyanobacteria. Bioresour. Technol. 177, 41–50.
Borges, F.C., Xie, Q., Min, M., Muniz, L.A.R., Farenzena, M., Trierweiler, J.O., Chen, P.,
Hu, Z., Ma, X., Li, L., 2016. The synergistic effect of co-pyrolysis of oil shale and
Ruan, R., 2014. Fast microwave-assisted pyrolysis of microalgae using
microalgae to produce syngas. J. Energy Inst. 89 (3), 447–455.
microwave absorbent and HZSM-5 catalyst. Bioresour. Technol. 166, 518–526.
Islam, M.A., Auta, M., Kabir, G., Hameed, B.H., 2016. A thermogravimetric analysis of
Branca, C., Di Blasi, C., 2004. Parallel- and series-reaction mechanisms of wood and
the combustion kinetics of karanja (Pongamia pinnata) fruit hulls char.
char combustion. Thermal Science 8, 51–63.
Bioresour. Technol. 200, 335–341.
Brennan, L., Owende, P., 2010. Biofuels from microalgae—a review of technologies
Janković, V., Schultz, D.M., 2017. Atmosfear: communicating the effects of climate
for production, processing, and extractions of biofuels and co-products. Renew.
change on extreme weather. Weather Climate Soc. 9 (1), 27–37.
Sustain. Energy Rev. 14 (2), 557–577.
Jena, U., Das, K.C., 2011. Comparative evaluation of thermochemical liquefaction
Broström, M., Nordin, A., Pommer, L., Branca, C., Di Blasi, C., 2012. Influence of
and pyrolysis for bio-oil production from microalgae. Energy Fuels 25 (11),
torrefaction on the devolatilization and oxidation kinetics of wood. J. Anal. Appl.
5472–5482.
Pyrol. 96, 100–109.
Kebelmann, K., Hornung, A., Karsten, U., Griffiths, G., 2013. Intermediate pyrolysis
Bui, H.-H., Tran, K.-Q., Chen, W.-H., 2016. Pyrolysis of microalgae residues – a kinetic
and product identification by TGA and Py-GC/MS of green microalgae and their
study. Bioresour. Technol. 199, 362–366.
extracted protein and lipid components. Biomass Bioenergy 49, 38–48.
Campanella, A., Harold, M.P., 2012. Fast pyrolysis of microalgae in a falling solids
Kim, S.-S., Ly, H.V., Choi, G.-H., Kim, J., Woo, H.C., 2012. Pyrolysis characteristics and
reactor: effects of process variables and zeolite catalysts. Biomass Bioenergy 46,
kinetics of the alga Saccharina japonica. Bioresour. Technol. 123, 445–451.
218–232.
Kim, S.-S., Ly, H.V., Kim, J., Choi, J.H., Woo, H.C., 2013. Thermogravimetric
Ceylan, S., Kazan, D., 2015. Pyrolysis kinetics and thermal characteristics of
characteristics and pyrolysis kinetics of Alga Sagarssum sp. biomass.
microalgae Nannochloropsis oculata and Tetraselmis sp. Bioresour. Technol.
Bioresour. Technol. 139, 242–248.
187, 1–5.
Kim, S.W., Koo, B.S., Lee, D.H., 2014. A comparative study of bio-oils from pyrolysis
Chaiwong, K., Kiatsiriroat, T., Vorayos, N., Thararax, C., 2013. Study of bio-oil and
of microalgae and oil seed waste in a fluidized bed. Bioresour. Technol. 162, 96–
bio-char production from algae by slow pyrolysis. Biomass Bioenergy 56, 600–
102.
606.
100 Q.-V. Bach, W.-H. Chen / Bioresource Technology 246 (2017) 88–100

Kirtania, K., Bhattacharya, S., 2012a. Application of the distributed activation energy Sharara, M.A., Holeman, N., Sadaka, S.S., Costello, T.A., 2014. Pyrolysis kinetics of
model to the kinetic study of pyrolysis of the fresh water algae Chlorococcum algal consortia grown using swine manure wastewater. Bioresour. Technol. 169,
humicola. Bioresour. Technol. 107, 476–481. 658–666.
Laird, D.A., Brown, R.C., Amonette, J.E., Lehmann, J., 2009. Review of the pyrolysis Shuping, Z., Yulong, W., Mingde, Y., Chun, L., Junmao, T., 2010. Pyrolysis
platform for coproducing bio-oil and biochar. Biofuels, Bioprod. Biorefin. 3 (5), characteristics and kinetics of the marine microalgae Dunaliella tertiolecta
547–562. using thermogravimetric analyzer. Bioresour. Technol. 101 (1), 359–365.
Lee, J.W., Hawkins, B., Day, D.M., Reicosky, D.C., 2010. Sustainability: the capacity of Singh, A., Olsen, S.I., 2011. A critical review of biochemical conversion, sustainability
smokeless biomass pyrolysis for energy production, global carbon capture and and life cycle assessment of algal biofuels. Appl. Energy 88 (10), 3548–3555.
sequestration. Energy Environ. Sci. 3 (11), 1695–1705. Sirajunnisa, A.R., Surendhiran, D., 2016. Algae – a quintessential and positive
Li, F., Srivatsa, S.C., Batchelor, W., Bhattacharya, S., 2017. A study on growth and resource of bioethanol production: a comprehensive review. Renew. Sustain.
pyrolysis characteristics of microalgae using thermogravimetric analysis- Energy Rev. 66, 248–267.
infrared spectroscopy and synchrotron fourier transform infrared Suali, E., Sarbatly, R., 2012. Conversion of microalgae to biofuel. Renew. Sustain.
spectroscopy. Bioresour. Technol. 229, 1–10. Energy Rev. 16 (6), 4316–4342.
López-González, D., Fernandez-Lopez, M., Valverde, J.L., Sanchez-Silva, L., 2014. Tang, Y., Ma, X., Lai, Z., 2011. Thermogravimetric analysis of the combustion of
Pyrolysis of three different types of microalgae: kinetic and evolved gas microalgae and microalgae blended with waste in N 2/O 2 and CO 2/O 2
analysis. Energy 73, 33–43. atmospheres. Bioresour. Technol. 102 (2), 1879–1885.
López-González, D., Puig-Gamero, M., Acién, F.G., García-Cuadra, F., Valverde, J.L., Tran, K.-Q., Bach, Q.-V., Trinh, T.T., Seisenbaeva, G., 2014. Non-isothermal pyrolysis
Sanchez-Silva, L., 2015. Energetic, economic and environmental assessment of of torrefied stump – a comparative kinetic evaluation. Appl. Energy 136, 759–
the pyrolysis and combustion of microalgae and their oils. Renew. Sustain. 766.
Energy Rev. 51, 1752–1770. Uggetti, E., Sialve, B., Trably, E., Steyer, J.P., 2014. Integrating microalgae production
Lü, J., Sheahan, C., Fu, P., 2011. Metabolic engineering of algae for fourth generation with anaerobic digestion: a biorefinery approach. Biofuels, Bioprod. Biorefin. 8
biofuels production. Energy Environ. Sci. 4 (7), 2451–2466. (4), 516–529.
Manyà, J.J., 2012. Pyrolysis for biochar purposes: a review to establish current Várhegyi, G., 2007. Aims and methods in non-isothermal reaction kinetics. J. Anal.
knowledge gaps and research needs. Environ. Sci. Technol. 46 (15), 7939–7954. Appl. Pyrol. 79 (1–2), 278–288.
Marcilla, A., Gómez-Siurana, A., Gomis, C., Chápuli, E., Catalá, M.C., Valdés, F.J., 2009. Várhegyi, G.B., Bobály, B.Z., Jakab, E., Chen, H., 2010. Thermogravimetric study of
Characterization of microalgal species through TGA/FTIR analysis: application biomass pyrolysis kinetics. A distributed activation energy model with
to nannochloropsis sp. Thermochim. Acta 484 (1–2), 41–47. prediction tests. Energy Fuels 25 (1), 24–32.
Mata, T.M., Martins, A.A., Caetano, N.S., 2010. Microalgae for biodiesel production Wang, K., Brown, R.C., Homsy, S., Martinez, L., Sidhu, S.S., 2013. Fast pyrolysis of
and other applications: a review. Renew. Sustain. Energy Rev. 14 (1), 217–232. microalgae remnants in a fluidized bed reactor for bio-oil and biochar
Miao, X., Wu, Q., 2004. High yield bio-oil production from fast pyrolysis by production. Bioresour. Technol. 127, 494–499.
metabolic controlling of Chlorella protothecoides. J. Biotechnol. 110 (1), 85–93. Wang, X., Zhao, B., Tang, X., Yang, X., 2015. Comparison of direct and indirect
Miao, X., Wu, Q., Yang, C., 2004. Fast pyrolysis of microalgae to produce renewable pyrolysis of micro-algae Isochrysis. Bioresour. Technol. 179, 58–62.
fuels. J. Anal. Appl. Pyrol. 71 (2), 855–863. Wang, X., Sheng, L., Yang, X., 2017. Pyrolysis characteristics and pathways of
Miura, K., 1995. A new and simple method to estimate f(E) and k0(E) in the protein, lipid and carbohydrate isolated from microalgae Nannochloropsis sp.
distributed activation energy model from three sets of experimental data. Bioresour. Technol. 229, 119–125.
Energy Fuels 9 (2), 302–307. White, J.E., Catallo, W.J., Legendre, B.L., 2011. Biomass pyrolysis kinetics: a
Nakanishi, A., Aikawa, S., Ho, S.-H., Chen, C.-Y., Chang, J.-S., Hasunuma, T., Kondo, A., comparative critical review with relevant agricultural residue case studies. J.
2014. Development of lipid productivities under different CO2 conditions of Anal. Appl. Pyrol. 91 (1), 1–33.
marine microalgae Chlamydomonas sp. JSC4. Bioresour. Technol. 152, 247–252. Wildschut, J., Melian-Cabrera, I., Heeres, H., 2010. Catalyst studies on the
Orfão, J.J.M., Antunes, F.J.A., Figueiredo, J.L., 1999. Pyrolysis kinetics of hydrotreatment of fast pyrolysis oil. Appl. Catal. B 99 (1), 298–306.
lignocellulosic materials—three independent reactions model. Fuel 78 (3), Wu, K., Liu, J., Wu, Y., Chen, Y., Li, Q., Xiao, X., Yang, M., 2014. Pyrolysis
349–358. characteristics and kinetics of aquatic biomass using thermogravimetric
Pan, P., Hu, C., Yang, W., Li, Y., Dong, L., Zhu, L., Tong, D., Qing, R., Fan, Y., 2010. The analyzer. Bioresour. Technol. 163, 18–25.
direct pyrolysis and catalytic pyrolysis of Nannochloropsis sp. residue for Wu, X., Wu, Y., Wu, K., Chen, Y., Hu, H., Yang, M., 2015. Study on pyrolytic kinetics
renewable bio-oils. Bioresour. Technol. 101 (12), 4593–4599. and behavior: the co-pyrolysis of microalgae and polypropylene. Bioresour.
Peng, W., Wu, Q., Tu, P., 2001. Pyrolytic characteristics of heterotrophic Chlorella Technol. 192, 522–528.
protothecoides for renewable bio-fuel production. J. Appl. Phycol. 13 (1), 5–12. Xie, Q., Addy, M., Liu, S., Zhang, B., Cheng, Y., Wan, Y., Li, Y., Liu, Y., Lin, X., Chen, P.,
Peters, J.F., Banks, S.W., Bridgwater, A.V., Dufour, J., 2017. A kinetic reaction model Ruan, R., 2015. Fast microwave-assisted catalytic co-pyrolysis of microalgae and
for biomass pyrolysis processes in Aspen Plus. Appl. Energy 188, 595–603. scum for bio-oil production. Fuel 160, 577–582.
Please, C.P., McGuinness, M.J., McElwain, D.L.S., 2003. Approximations to the Xiu, S., Shahbazi, A., 2012. Bio-oil production and upgrading research: a review.
distributed activation energy model for the pyrolysis of coal. Combust. Flame Renew. Sustain. Energy Rev. 16 (7), 4406–4414.
133 (1–2), 107–117. Xu, G., Lv, Y., Sun, J., Shao, H., Wei, L., 2012. Recent advances in biochar applications
Raheem, A., Azlina, W.W., Yap, Y.T., Danquah, M.K., Harun, R., 2015. in agricultural soils: benefits and environmental implications. CLEAN – Soil Air
Thermochemical conversion of microalgal biomass for biofuel production. Water 40 (10), 1093–1098.
Renew. Sustain. Energy Rev. 49, 990–999. Yahiaoui, M., Hadoun, H., Toumert, I., Hassani, A., 2015. Determination of kinetic
Razeghifard, R., 2013. Algal biofuels. Photosynth. Res. 117 (1), 207–219. parameters of Phlomis bovei de Noé using thermogravimetric analysis.
Rizzo, A.M., Prussi, M., Bettucci, L., Libelli, I.M., Chiaramonti, D., 2013. Bioresour. Technol. 196, 441–447.
Characterization of microalga Chlorella as a fuel and its thermogravimetric Yang, H., Yan, R., Chen, H., Lee, D.H., Zheng, C., 2007. Characteristics of
behavior. Appl. Energy 102, 24–31. hemicellulose, cellulose and lignin pyrolysis. Fuel 86 (12), 1781–1788.
Saldarriaga, J.F., Aguado, R., Pablos, A., Amutio, M., Olazar, M., Bilbao, J., 2015. Fast Yang, X., Zhang, R., Fu, J., Geng, S., Cheng, J.J., Sun, Y., 2014. Pyrolysis kinetic and
characterization of biomass fuels by thermogravimetric analysis (TGA). Fuel product analysis of different microalgal biomass by distributed activation
140, 744–751. energy model and pyrolysis–gas chromatography–mass spectrometry.
Sambusiti, C., Bellucci, M., Zabaniotou, A., Beneduce, L., Monlau, F., 2015. Algae as Bioresour. Technol. 163, 335–342.
promising feedstocks for fermentative biohydrogen production according to a Zhao, B., Wang, X., Yang, X., 2015. Co-pyrolysis characteristics of microalgae
biorefinery approach: a comprehensive review. Renew. Sustain. Energy Rev. 44, Isochrysis and Chlorella: kinetics, biocrude yield and interaction. Bioresour.
20–36. Technol. 198, 332–339.
Sanchez-Silva, L., López-González, D., Garcia-Minguillan, A.M., Valverde, J.L., 2013. Zhu, L., 2015. Biorefinery as a promising approach to promote microalgae industry:
Pyrolysis, combustion and gasification characteristics of Nannochloropsis an innovative framework. Renew. Sustain. Energy Rev. 41, 1376–1384.
gaditana microalgae. Bioresour. Technol. 130, 321–331.

You might also like