You are on page 1of 41

C H A P T E R

1
Biofuels: Introduction
Roger Ruan†, Yaning Zhang*,†, Paul Chen†, Shiyu Liu†,
Liangliang Fan‡, Nan Zhou†, Kuan Ding†, Peng Peng†, Min
Addy†, Yanling Cheng†, Erik Anderson†, Yunpu Wang‡,
Yuhuan Liu‡, Hanwu Lei§, Bingxi Li*
*
School of Energy Science and Engineering, Harbin Institute of Technology, Harbin, China

Center for Biorefining, Department of Bioproducts and Biosystems Engineering, University of
Minnesota, St Paul, MN, United States ‡State Key Laboratory of Food Science and Technology,
Nanchang University, Nanchang, China §Department of Biological Systems Engineering,
Washington State University, Richland, WA, United States

1.1 INTRODUCTION

Biofuels have been serving mankind for a very long time ever since we started burning
firewood for heating and cooking. In this regard, biofuels even have a longer history than
humanity’s history and civilization because before we humans realized and used the fire,
biofuels have already existed, for example, in the form of wood and grass.
As human civilization went on, biofuels played more and more important roles in the peo-
ple’s lives, for example, methane gas (biomethane or biogas) for preparing warm bath water
in the l0th century BC [1], ethanol (bioethanol) in an internal combustion engine in 1826 [2],
Rudolph diesel (biodiesel) in Sadi Carnot engines in 1878 [3], peanut oil in the first trucks and
buses in 1896 [4], etc. Applications of biofuels in people’s lives have been detailed in many
publications [5–8]. These applications dramatically changed and improved the people’s life
all over the word.
As compared with fossil fuels which mainly include coal, oil, and natural gas, biofuels gen-
erally have the following advantages: (a) biofuels are renewable [9], (b) biofuels and resources
are abundantly available [10], (c) they are CO2/GHG (greenhouse gas) neutral [11], (d) they
result in negligible or zero SOx emissions [12], (e) they result in lower NOx emissions [13], (f )
they are environmental friendly fuels [14], (g) they can be produced locally [15], (h) they are

Biofuels: Alternative Feedstocks and Conversion Processes for the


Production of Liquid and Gaseous biofuels 3 # 2019 Elsevier Inc. All rights reserved.
https://doi.org/10.1016/B978-0-12-816856-1.00001-4
4 1. BIOFUELS: INTRODUCTION

biodegradable [16], (i) they are sustainable [17], (j) the production processes and/or farms of
biofuels are generally safe [18], etc.
The biofuels also have the following benefits: (a) they utilize local energy resources [19],
(b) they boost rural agricultural development and investment [20], (c) they enhance supply
diversity [20], (d) they reduce energy dependence on imports [18, 19], (e) they create jobs
[18, 20], (f ) they promote rural economic development and improve local livelihood [21], etc.
Therefore, many countries have projected numerous correlated targets or policies to
promote and develop biofuel industry. These include documents issued in the United States
(USA), the European Union (EU), Brazil, China, India, and Thailand (Table 1.1).

1.2 BIOFUEL

1.2.1 Definition
In the early 19th century, the term “fossil fuel” was used to refer to fuels such as coal, oil,
and natural gas which were formed from the remains of plants and animals buried beneath
the Earth’s surface for millions of years [23]. In the 1970s, a new word, “biofuel,” appeared
and began to be used.
Up to now, there are many different definitions for the term “biofuel” or “bio-fuel.” Some
statements are as follows:
(a) Biofuel is fuel that is produced from, or by, recently living organisms [18].
(b) Biofuel is any fuel that is in some way derived from biomass, including firewood,
methane, and petroleum [24].
(c) A biofuel is a fuel made from plants or other biological materials [25].
(d) The term biofuel or biorenewable fuel is referred to as solid, liquid, or gaseous fuels that
are predominantly produced from biomass [26].

TABLE 1.1 Policies Set by Main Global Biofuels Producers


Country Target or Mandate References
EU Mandate: minimum of 10% of transport fuel from renewable fuels by 2020 [6]
United Mandate: 36 billion gallons of biofuel by 2022 [6]
States
Brazil Mandate: biodiesel use set at 10% by 2020 [6]
India Indicative: 20% blending for both ethanol and biodiesel by 2017 [6]
3
China Target: solid biofuels (10,000 t/year), biogas (billion m /year), nonfood bioethanol [22]
(10,000 t/year), biodiesel (10,000 t/year) by 2020
China Target: 12.7 Bnl ethanol and 2.3 Bnl biodiesel consumption in 2020 [6]
15% of fuel consumption to be nonfossil fuel by 2020
Thailand Ethanol: E20 mandatory since 2008 [6]
Biodiesel: B2 mandatory since 2008 and B5 since 2012

I. GENERAL
1.2 BIOFUEL 5
(e) Biofuels refer to plant biomass and the refined products to be combusted for energy (heat
and light) [1].
(f ) Biofuels may be defined as bio-based products found or extracted from nature, such as
wood, bagasse or peat, or chemically transformed from biomass to form products such as
charcoal, biooil, ethanol, and biogas. It can also be defined as biomass that is used as fuel [4].
(g) Biofuels are fuels derived from biomass or waste feedstock [20].
(h) Biofuels are fuels produced directly or indirectly from biomass such as fuelwood,
charcoal, bioethanol, biodiesel, biogas (methane), or biohydrogen [27, 28].
(i) Biofuels refer to liquid and gaseous fuels produced from biomass—organic matter
derived from plants or animals [29].
(j) Biofuels are renewable liquid fuels coming from biological raw material and have been
proved to be good substitutes for oil in the transportation sector [30].
(k) Biofuels are liquid or gaseous fuels for the transport sector that are predominantly
produced from a variety of bio-feedstocks. Bio-feedstock or biomass refers to all the
vegetable matters that can be obtained from photosynthesis [14].
(l) Biofuels are liquid or gaseous fuels for the transport sector that are produced from
renewable sources such as vegetable oil and biomass [31].
(m) Biofuels refer to those liquid or gaseous fuels, used for the transport sector, that are
predominantly produced from biomass [19, 32].
(n) Biofuels are liquid fuels derived from biomass, including biodiesel from fats and oils and
bioethanol from sugars, starches, and lignocellulosic materials [33].
(o) Biofuels are liquid transportation fuels derived from plants or other biological materials
[34], etc.
Based on the above definitions, “biofuel” can be defined as any fuel that is made by or from
a renewable living organism.

1.2.2 Implication
Based on its definition, the term “biofuel” or “bio-fuel” has at least three implications:
(a) A biofuel is also a fuel which can be used to generate energy, heat, light, or power.
(b) A biofuel is living organisms or from living organisms: the prefix “bio” means “life” or
“living organism” (“bio” is borrowed from the Greek “bios,” which means “mode of
life”), and this also makes it different from the other fuels like C or H2, which are not living
organisms but fuels.
(c) Biofuel is renewable which means the living organism can regenerate in a relatively short
period of time rather than a long time, for example, over millions of years (this
significantly differentiates biofuels from fossil fuels).
There are also some other implications which are as follows:
(a) A biofuel can be or from a plant part, animal waste, manure, sludge, etc.
(b) A biofuel can be in the form of solid, liquid, or gas.
(c) A biofuel can be converted to another biofuel (generally, the original or raw biofuel is
defined as a feedstock), etc.

I. GENERAL
6 1. BIOFUELS: INTRODUCTION

1.3 CLASSIFICATION

Similar to the various definitions of “biofuel,” there are also different classifications for
biofuels. According to the commercialization status of biofuel technologies, biofuels can be
classified into two groups: conventional biofuels and advanced biofuels (Fig. 1.1). The com-
mercialization degree is ranked as follows: commercial > early commercial > demonstration
> research. The conventional biofuels mainly include sugar- and starch-based bioethanol,
transesterification-based biodiesel, and anaerobic digestion biomethane (biogas). For the
advanced biofuels, hydrotreated vegetable oil is at an early commercial stage; cellulosic
(lignocellulosic) bioethanol, biomass-to-liquid (BtL) biodiesel, microalgae biodiesel, and
biohydrogen are mainly at the demonstration and research stages. It should be noted that
the advanced biodiesels are not widely available at present, but they may become fully
commercialized in the (near) future.
According to the feedstock sources and generation technologies, some researchers have
classified biofuels into two groups: first-generation biofuels and second-generation bio-
fuels [35, 36]. Fig. 1.2 shows the classifications of these two groups. The first-generation
biofuels mainly include bioethanol, biodiesel (bio-esters), and biogas [35, 37]. The liquid biofuels
mainly include saturated and unsaturated edible plant oils from the oilseeds of canola, corn,
sunflower, and soybean. Oils from the fruits of coconut, olive, and palm are also included
[36]. Morone and Cottoni [37] stated that the solid biofuels of firewood, dried manure, and
agricultural waste are also included in the first-generation biofuels. The second-generation
biofuels mainly include lipid-based biofuels which are produced from waste vegetable oils,
animal fats, insects, and oleaginous microorganisms [36].
The first-generation biofuels are commercialized today, whereas the second-generation
biofuels are still under development at present [35]. The first-generation biofuels and
second-generation biofuels are therefore similar to conventional biofuels and advanced
biofuels (Fig. 1.1), respectively.

Conventional biofuels Advanced biofuels

Commercial Early commercial Demonstration Research

Bioethanol Sugar and starch Cellulosic

Biodiesel Transesterification HVO BtL Microalgae

Biomethane Biogas Biosyngas


Biohydrogen Biohydrogen

FIG. 1.1 Biofuels classified according to the commercialization status. Modified from IEA, Biofuel Roadmap, http://
www.ieabioenergy.com/wp-content/uploads/2013/10/IEA-Biofuel-Roadmap.pdf, 2011.

First generation biofuels bloethanol, bIodiesel, biogas, etc.


Biofuels
Second generation biofuels vegetable oils, animal fats, etc.

FIG. 1.2 Biofuels classified into two generations.

I. GENERAL
1.3 CLASSIFICATION 7

First generation biofuels from food crops

Biofuels Second generation biofuels from non-food feedstocks

Third generation biofuels from aquatic cultivated feedstocks

FIG. 1.3 Biofuels classified into three generations.

Many researchers expanded the two generations of biofuels (first-generation biofuels and
second-generation biofuels) to three generations of biofuels: first-generation biofuels, second-
generation biofuels, and third-generation biofuels [7, 14, 38–43]. Fig. 1.3 outlines the three
groups of biofuels.
The first-generation biofuels mainly include the biofuels of biodiesel, bioalcohols, vegeta-
ble oil, biogas, and biosyngas produced from food crops like corn, wheat, and soybeans [41].
They are mainly produced from sugar, starch, and vegetable oil [43]. Gupta et al. [39] also
included wood, grains, straw, switchgrass, perennial rye grass, grass press cake, grass cut-
tings, charcoal, domestic refuse, and dried manure in the first-generation biofuels.
The second-generation biofuels mainly include the biofuels of bioethanol, biodiesel,
dimethyl-ether, biosyngas FT (Fischer-Tropsch) biodiesel, BtL (biomass to liquid) biodiesel,
etc. These biofuels are mainly produced from nonfood feedstocks (mainly lignocellulosic bio-
mass) such as grass, wood, cereal straw, sugarcane bagasse, forest residues, dedicated energy
crops (purpose-grown vegetative grasses and short rotation forests), as well as municipal
solid wastes [7, 41, 44].
The third-generation biofuels mainly include the biofuels of biomethane, biodiesel,
bioethanol, biobutanol, vegetable oil gasoline, jet fuels, and aviation fuels which are mainly
derived from aquatic cultivated feedstocks like algae or cyanobacteria [14, 38, 43–46].
Algayyim et al. [47] included the solid biofuels of grass, wood, and wood chips in the pri-
mary biofuels which can be directly used for heating or cooking without modification. The
secondary biofuels include the three generations of biofuels (first-generation biofuels,
second-generation biofuels, and third-generation biofuels). Then, biofuels are classified into
two broad groups as shown in Fig. 1.4.
Noraini et al. [48] separately defined natural biofuels which are generally derived from
organic sources such as firewood, plants, vegetables, animal waste, and landfill gas. These
natural biofuels are commonly used for cooking, heating, brick kiln, or electricity production.
The first-generation biofuels are derived from edible feedstocks like wheat, palm, corn,
maize, soybean, sugarcane, rapeseed, oil crops, sugar beet, etc. [48]. Lignocellulosic feed-
stocks such as miscanthus, jatropha, sterculia foetida, ceiba pentandra, switchgrass, and

Primary biofuels Without modification grass, wood, etc.

Biofuels First generation biofuels from grains, sugars

Secondary biofuels Second generation biofuels from lignocellulosic

Third generation biofuels from algae, microbes

FIG. 1.4 Biofuels classified into primary biofuels and secondary biofuels.

I. GENERAL
8 1. BIOFUELS: INTRODUCTION

Natural biofuels firewood, animal waste, landfill gas

First generation biofuels from edible feedstocks


Biofuels
Second generation biofuels from lignocellulosic feedstocks

Third generation biofuels from algae feedstocks

FIG. 1.5 Biofuels classified into four groups according to Noraini et al. [48].

poplar are included in the second-generation biofuels. The third-generation biofuels are
mainly derived from algae feedstocks. The classification is outlined in Fig. 1.5.
Hao et al. [49] combined oil-bearing crops (jatropha) and nonfood crops (cassava), and de-
fined the produced biofuels as 1.5-generation biofuels. The biofuels were then classified into
four groups: first-generation (1G) biofuels, 1.5-generation (1.5G) biofuels, second-generation
(2G) biofuels, and third-generation (3G) biofuels (Fig. 1.6).
Most of the researchers categorized biofuels into another four generations biofuels: first-
generation biofuels, second-generation biofuels, third-generation biofuels, and fourth-
generation biofuels [3, 6, 26, 39, 50–56].
The first-generation biofuels mainly include bioalcohols, vegetable oil, biodiesel,
biomethanol, biosyngas, and biogas. The second-generation biofuels mainly include
bioalcohols, biooil, biodiesel, bioDMF (dimethylfuran), biomethanol, bioFT diesel, and
biohydrogen. The third-generation biofuels mainly include bioethanol, vegetable oil, biodie-
sel, biomethanol, and jet fuels. The fourth-generation biofuels mainly include green diesel,
biogasoline, and green aviation fuels. Fig. 1.7 shows the outline of the four groups and
Table 1.2 presents the biofuel examples and feedstocks used.
Regarding these four generations biofuels, the advantages and disadvantages of each type
are summarized in Table 1.3.
Based on the above classifications, biofuels can be classified into the following five groups:
zeroth-generation biofuels, first-generation biofuels, second-generation biofuels, third-

Food corn, wheat 1G biofuels


Agricultural crops
Non-food cassava
1.5G biofuels
Plantation Oil-bearing crops jatropha

Cellulosic crops fast growing grass

Agricultural residues corn cob, straw 2G biofuels

Biomass Non-plantation Forestry residues forest residues

Waste oil waste cooking oil

Algae algae 3G biofuels

FIG. 1.6 Biofuels classified into four groups. Modified from H. Hao, Z. Liu, F. Zhao, J. Ren, S. Chang, K. Rong, J. Du, Biofuel
for vehicle use in China: current status, future potential and policy implications, Renew. Sust. Energ. Rev. 82 (2018) 645–653.

I. GENERAL
1.3 CLASSIFICATION 9

First generatioin biofuels bioalcohols, vegetable oil, biodiesel, biomethanol, biosyngas, biogas

Second generatioin biofuels bioalcohols, biooil, biodiesel, bioDMF, biomethanol, bioFT diesel, biohydrogen
Biofuels
Third generatioin biofuels bioethanol, vegetable oil biodiesel, biomethanol, jet fuels

Fourth generatioin biofuels green biodiesel, biogasoline, green aviation fuels

FIG. 1.7 Biofuels classified into the widely accepted four groups.

TABLE 1.2 The Four Generation Biofuels and the Feedstocks Used [3, 6, 26, 39, 53–56]
Biofuels Examples Feedstocks
First Bioalcohols, vegetable oil, biodiesel, Sugar, starch, animal fats, soybean, soya, rapeseed,
generation biomethanol, biosyngas, biogas mustard, sunflower, maize, sugarcane, sugar beet,
sorghum, potato, palm oil, coconut, canola, plant,
cassava, castor, jatropha, sewage waste
Second Bioalcohols, biooil, biodiesel, BioDMF, Nonfood crops, wheat straw, corn, wood, switchgrass,
generation biomethanol, bioFischer-Tropsch diesel, cereal straw, sugarcane bagasse, reed canary grass,
biohydrogen forest residues, energy crops, municipal solid wastes,
alfalfa, agave, jatropha
Third Bioethanol, vegetable oil, biodiesel, Microbial species, algae, yeast, fungi, cyanobacteria
generation biomethanol, jet fuels

Fourth Green diesel, biogasoline, green aviation fuel Vegetable oil, biodiesel
generation

generation biofuels, and fourth-generation biofuels (Fig. 1.8). The first-generation biofuels,
second-generation biofuels, third-generation biofuels, and fourth-generation biofuels are
similar to the ones detailed previously. The zeroth-generation biofuels are mainly the natural
biofuels or raw feedstocks which can be directly used without pretreatment, processing, or
modifications.

TABLE 1.3 Advantages and Disadvantages of the Four Generation Biofuels [6, 54, 57]
Biofuel Advantages Disadvantages
First (1) Biodegradable (1) Competition of land use
generation (2) Energy security (2) Blending with conventional fuel
(3) Feedstock can be easily produced by already (3) Highest carbon footprint compared with
existing infrastructure and technology other generations of biofuel
(4) Environmental and social benefits (4) Requires large amount of inputs in terms of
(5) Feedstocks available at large quantities fertilizer, water and land area thereby
reducing net energy ratio
(5) Contributes to higher food prices owing to
competition with food
(6) Might potentially have negative impacts on
biodiversity

Continued

I. GENERAL
10 1. BIOFUELS: INTRODUCTION

TABLE 1.3 Advantages and Disadvantages of the Four Generation Biofuels [6, 54, 57]—cont’d
Biofuel Advantages Disadvantages
Second (1) No competition with food (1) Even though requires less compared to first-
generation (2) Use of whole plant instead of only seeds or generation biofuels, the land required for
grains, and use of residues means more energy production of second-generation feedstock is
produced per hectare of land substantial
(3) Marginal lands can be used for planting of (2) The use of agriculture and forest residue
advanced feedstock such as Jatropha sp. degrades soil quality and also induces soil
(4) Higher yield and lower land requirement erosion
(5) Available feedstocks in large quantities (3) Complex processes are required
(6) Feedstock can be easily produced by already (4) Low conversion as compared with petroleum
existing infrastructure and technology fuel
(7) Low cost for feedstock (5) Conversion technologies are under
(8) Energy security development
(9) Production of high-value added products (6) Lack of technological and research
(10) Close to meeting the claimed environmental breakthrough
benefits (7) Lack of efficient technologies for commercial
applications

Third (1) No food or land competition (1) Difficult to harvest and process
generation (2) Produces more energy per acre than (2) High processing cost
conventional crops (3) Lack of technological and research
(3) Algae can be grown using land and water breakthrough
unsuitable for food production (4) Not yet commercially feasible
(4) High oil yield (5) Production technology is under development
(5) No toxic content (6) Require new technologies from the
(6) Energy security production of feedstock to processing into
(7) Bioengineered algae are renewable final biofuel product
(8) Low and sometimes no cost for feedstock
(9) Improves performance of first- and second-
generation biofuels when employed in
integrated biofuels

Fourth (1) Fourth-generation biofuel is argued to be (1) Lack of study on its practical performance in
generation carbon negative rather than simply carbon terms of technical and economic aspects
neutral, as it “locks” away more carbon than it (2) High cost
produces (3) Still in research and development stage
(2) Synthetic raw materials to produce biofuels is a (4) Requires new technologies from the
possibility production of feedstock to processing into
(3) Energy security final biofuel product

Zero generatioin biofuels natural biofuels

First generatioin biofuels bioalcohols, vegetable oil, biodiesel, biomethanol, biosyngas, biogas

Biofuels Second generatioin biofuels bioalcohols, biooil, biodiesel, bioDMF, biomethanol, bioFT diesel, biohydrogen

Third generatioin biofuels bioethanol, vegetable oil biodiesel, biomethanol, jet fuels

Fourth generatioin biofuels green biodiesel, biogasoline, green aviation fuels

FIG. 1.8 Biofuels classified into five generations.

I. GENERAL
1.4 BIOFUEL FEEDSTOCK 11

1.4 BIOFUEL FEEDSTOCK

Biofuels are mainly generated from biomass. There are also many classifications for bio-
mass [6, 58–61]. As for the feedstocks used for the different generations of biofuels
(Table 1.2), the same feedstock may belong to different groups, for example, sugarcane be-
longs to the first-generation biofuel feedstock and also to the second-generation biofuel feed-
stock. In this section, some typical biofuel feedstocks are detailed, without sticking to the
classifications of biomass or biofuels. The typical biofuel feedstocks include: (a) oil crops,
(b) lignocellulosic biomass, (c) solid waste, and (d) algae. Some other biofuel feedstocks
are also referred to in other publications, for example, bacterium, yeast, fungi, etc. [62].

1.4.1 Oil Crops


Oil crops mainly include sunflower,soybean, corn, rape seed, cotton seed, oil palm, palm
kernel, palm fruit, coconut, peanut, canola, copra, castor, polanga, jatropha, sesame, etc. As
the name tells, an oil crop contains oil. Table 1.4 illustrates the oil contents of differentoil
crops. It is observed that oil contents vary signincantly among different oil crops, for example,
coconut and polanga have much higher oil contents (65–75 wt%) than cotton seed (13 wt%). It
is also observed that the same oil crop may have quite different oil contents, for example, the
oil content of jatropha significantly varies between 20 and 60 wt%. It may be due to the dif-
ferences in the plant conditions, conversion methods, and processing equipment. Generally,
coconut, polanga, peanut, copra, jatropha, sesame, castor, linseed, corn, camelina, groundnut
kernel, canola, and sunflower are good oil crops which can be used as feedstocks for biofuel
(biooil or biodiesel) production.

1.4.2 Lignocellulosic Biomass


Lignocellulosic biomass mainly includes: (a) agricultural residues, such as rice straw, rice
husk, wheat straw, sorghum straw, corn stover, sugarcane bagasse, etc., (b) forest residues,
such as woods, woodchips, wood branches, wood sawdust, etc., and (c) energy crops, such as
switchgrass, miscanthus, energycane, grass, etc. Lignocellulosic biomass is mainly composed
of cellulose, hemicellulose, and lignin. Table 1.5 presents the contents of cellulose, hemicel-
lulose, and lignin for some typical lignocellulosic biomass. It is observed that the contents of
cellulose, hemicellulose, and lignin are generally 30.42–49.85 wt%, 18.00–35.00 wt%, and
7.00–36.02 wt%, respectively, for the agricultural residues. The contents of cellulose, hemicel-
lulose, and lignin are generally 23.70–59.70 wt%, 13.00–39.00 wt%, and 18.10–34.00 wt%,
respectively, for the forest residues. The chemical contents of cellulose, hemicellulose, and
lignin are generally 28.00–49.00 wt%, 15.00–32.17 wt%, and 4.00–25.94 wt%, respectively,
for the energy crops. Generally, agricultural residues (7.00–36.02 wt%) and forest residues
(18.10–34.00 wt%) have higher lignin contents than energy crops (4.00–25.94 wt%). Lignin
is generally difficult to treat or convert and it may make the pretreatment of lignocellulosic
biomass more difficult. A lower lignin content is therefore preferred. It is also observed
that the differences within a group are generally greater than the differences among the
groups, for example, the differences in the cellulose contents between two fruit bunches

I. GENERAL
12 1. BIOFUELS: INTRODUCTION

TABLE 1.4 Oil Contents of Oil Crops


Species Oil Fraction (wt%) References
Coconut 65–75 [42]

Polanga 65–75 [42]


Peanut 70 [42]
Copra 62 [63]
Jatropha 50–60 [64]
Jatropha 20–60 [42]
Sesame 50 [63]

Castor 45–50 [64]


Castor 48 [42]
Linseed 35–45 [64]
Corn 44 [42]
Camelina 42 [42]

Groundnut kernel 42 [63]


Canola 41 [42]
Sunflower 40 [42]
Sunflower 32 [63]
Mahua 35–40 [64]
Pongamia (Karanja) 30–40 [64]

Karanj 25–40 [42]


Rape seed 37 [63]
Oil palm 36 [42]
Palm kernel 36 [63]
Mustard 35 [63]
Neem 20–30 [64]

Safflower 20.1 [42]


Palm fruit 20 [63]
Soybean 18 [42]
Soybean 14 [63]
Cotton seed 13 [63]
Sal 10–12 [64]

I. GENERAL
1.4 BIOFUEL FEEDSTOCK 13
TABLE 1.5 Chemical Compositions of Lignocellulosic Biomass
Content (wt%)
Species Cellulose Hemicellulose Lignin References

AGRICULTURAL RESIDUES

Rice straw 35.00 25.00 12.00 [65]


Rice straw 40.00 18.00 7.00 [61]
Rice straw 42.00 20.00 26.00 [66]
Rice husk 30.42 28.03 36.02 [67]
Wheat straw 40.00 28.00 16.00 [61]
Sorghum straw 44.00 35.00 15.00 [61]

Barley straw 43.00 30.00 7.00 [61]


Corn stover 37.00 27.00 18.00 [68]
Corn stover 46.00 35.00 19.00 [61]
Corn stover 48.30 31.37 14.46 [69]
Bagasse 33.00 30.00 29.00 [61]
Sugarcane bagasse 49.85 28.36 14.92 [70]

Sugarcane bagasse 46.55 27.40 20.61 [67]


Sugarcane peel 41.11 26.40 24.31 [67]
FOREST RESIDUES
Aspen wood 53.00 27.00 19.00 [71]
Sawdust 48.00 15.00 32.00 [65]
Fruit bunch 59.70 22.10 18.10 [72]

Fruit bunch 23.70 21.60 29.20 [73]


Willow chips 37.20 36.00 18.70 [74]
Black locust 42.00 18.00 27.00 [61]
Hybrid poplar 45.00 19.00 26.00 [61]
Eucalyptus 50.00 13.00 28.00 [61]
Spruce 43.00 26.00 29.00 [61]

Pine 45.00 20.00 29.00 [61]


Softwood 40–45 30 26–34 [75]
Hardwood 40–50 23–39 20–30 [75]
Continued

I. GENERAL
14 1. BIOFUELS: INTRODUCTION

TABLE 1.5 Chemical Compositions of Lignocellulosic Biomass—cont’d


Content (wt%)
Species Cellulose Hemicellulose Lignin References

ENERGY CROPS

Switchgrass 37.66 32.17 25.94 [76]


Miscanthus 40.71 29.04 24.96 [76]
Hybrid pennisetum 41.52 29.65 23.81 [76]
Triarrhena lutarioriparia 40.77 29.82 25.14 [76]
Energycane leaf 35–39 26–29 5–7 [77]
Energycane stem 28–34 15–24 5–6 [77]

Grass leaf 31–38 24–30 4–5 [77]


Grass stem 39–49 17–23 7–13 [77]

(23.70–59.70 wt%) are much higher than the differences in the cellulose contents of the three
groups (30.42–49.85 wt%, 23.70–59.70 wt%, and 28.00–49.00 wt% for agricultural residues,
forest residues, and energy crops, respectively).

1.4.3 Solid Wastes


Solid wastes mainly inlclude municipal solid waste (paper, plastics, etc.), wastewater
sludge, food waste, animal manure, etc. Table 1.6 presents the chemical compositions of some
solid wastes. It is observed that there are significant differences among the chemical compo-
sitions of solid wastes, for example, plastics has a high content of cellulose (65 wt%) whereas
cattle manure has a low content of cellulose (2.7 wt%).
Generally, solid wastes are rich in moisture or water, for example, wastewater sludge, animal
manure, fish waste, etc. This high moisture content (may be higher than 80wt%) renders the fuel
(solid waste) a low-energy content (heating value). This makes it difficult to treat or convert them
through certain technologies, for example, combustion, gasification, etc. Yet, on the other hand, it
allows easy handling through other technologies, for example, digestion, extraction, etc.

TABLE 1.6 Chemical Compositions of Solid Waste [61]


Species Cellulose (wt%) Hemicellulose (wt%) Lignin (wt%)
Processed paper 47 25 12
Plastics 65 15 7.5
Food waste 45 5.3 13

Poultry waste 11 16 4
Cattle manure 2.7 2.3 4.5

I. GENERAL
1.4 BIOFUEL FEEDSTOCK 15

1.4.4 Algae
The potential of algae as a feedstock for biofuel production is attractive due to the
following desirable attributes: (a) algae are fast-growing microorganisms, (b) they are
promising nonfood resources, (c) they have simple cellular structure, (d) they have a
lipid-rich composition, (e) they can thrive in salty or wastewater, (f ) they tolerate marginal
lands (e.g., desert, arid, and semi-arid land), (g) they use carbon dioxide as the carbon
source, (h) they are nontoxic (no sulfur content), and (i) they are highly biodegradable
[14, 35, 46, 78].
Algae are photosynthetic aquatic organisms. The term “algae” refers to a polyphyletic, ar-
tificial assemblage of organisms [14]. If it is specifically used to refer to eukaryotic organisms,
photosynthetic bacteria are excluded. Algae are very diverse with over 40,000 species already
identified and many more not identified [14, 35]. Algae can be classified into two broad
groups of macroalgae (seaweeds, multicellular, or filamentous) and microalgae (unicellular,
microphytes, or phytoplankton) [14, 79]. As the names tell, macroalgae can reach sizes of up to
60 m in length whereas microalgae are always smaller than 0.4 mm [18].
Both macroalgae and microalgae can be subclassified into subgroups according to their
colors (pigmentation). Macroalgae can be subclassified into three major groups: Rhodophyta
(red), Chlorophyta (green), and Phaeophyta (brown) [42, 79, 80]. Microalgae can be
subdivided into four classes of Bacillariophyceae (diatom), Chlorophyceae (green),
Cyanophyceae (blue), and Chrysophyceae (golden) [81].
Both macroalgae and microalgae contain three main components: protein, carbohydrate,
and lipid. The chemical compositions of various macroalgae and microalgae are presented
in Table 1.7. It is observed that the chemical compositions of protein, carbohydrate, and lipid
are 5.06–20.93 wt%, 11.60–56.25 wt%, and 6.99–15.70 wt% for macroalgae and 6.00–71.00 wt%,
4.00–64.00 wt%, and 1.90–40.00 wt% for microalgae, respectively. Generally, microalgae have
higher contents of carbohydrate (6.00–1.00 wt%) and lipid (1.90–40.00 wt%) than macroalgae
(5.06–20.93 wt% and 6.99–15.70 wt%, respectively).
As compared with microalgae, macroalgae (or seaweeds) have higher volumetric pro-
duction rates (biomass per volume per time) and result in greater biomass densities [18].
Because macroalgae accumulate lipids as storage materials when they are stressed under
low growth rate conditions, they are therefore good feedstocks which can be used for
food production, hydrocolloid extraction, and fertilizer production [18, 80]. However,
the low digestibility of macroalgae cell wall makes it difficult to use macroalgae for biofuel
production [18], and production of biofuels from macroalgae biomass receives less
attention [80].
On the other hand, microalgae (e.g., cyanobacteria) store their lipids in thylakoid
membranes under the conditions of high photosynthetic rate and high growth rate. Gen-
erally, they are better than macroalgae for lipid production and are also better suited for
genetic improvement [80]. Also, the high carbohydrate content of microalgae (as illus-
trated in Table 1.7) can be used as carbon sources for bioethanol production [82], and
the high lipid content (as illustrated in Table 1.7) can be used as raw material for biodie-
sel production [14]. Consequently, biofuel production from microalgae is widely studied
and applied [45, 83].

I. GENERAL
16 1. BIOFUELS: INTRODUCTION

TABLE 1.7 Chemical Compositions of Macroalgae and Microalgae (wt% Dry Weight) [14]
Algae Protein Carbohydrate Lipid

MACROALGAE
Hypnea valentiae 11.8–12.6 11.8–13.0 9.6–11.6
Acanthophora spicifera 12.0–13.2 11.6–13.2 10.0–12.0

Laurencia papillosa 11.8–12.9 12.0–13.3 8.9–10.8


Ulva lactuca 11.4–12.6 11.6–13.2 9.6–11.4
Caulerpa racemosa 11.8–12.5 16.0 9.0–10.5
Ulva reticulate 12.83 16.88 8.50
Enteromorpha compressa 7.26 24.75 11.45
Chaetomorpha aerea 10.13 31.50 8.50

Chaetomorpha antennina 10.13 27.00 11.45


Chaetomorpha linoides 9.45 27.00 12.00
Cladophora fascicularis 15.53 49.50 15.70
Microdictyon agardhianum 20.93 27.00 9.40
Boergesenia forbesii 7.43 21.38 11.42
Valoniopsis pachynema 8.78 31.50 9.09

Dictyosphaeria cavernosa 6.00 42.75 10.51


Caulerpa cupressoides 7.43 51.75 10.97
Caulerpa peltata 6.41 45.00 11.42
Caulerpa laetevirens 8.78 56.25 8.80
Caulerpa racemosa 8.78 33.75 10.63
Caulerpa fergusonii 7.76 23.63 7.15

Caulerpa sertularioides 9.11 49.50 6.99


Halimeda macroloba 5.40 32.63 9.89
Codium adhaerens 7.26 40.50 7.40
Codium decorticatum 6.08 50.63 9.00
Codium tomentosum 5.06 29.25 7.15
MICROALGAE

Scenedesmus obliquus 50–56 10–17 12–14


Scenedesmus quadricauda 47 – 1.9
Scenedesmus dimorphus 8–18 21–52 16–40
Chlamydomonas rheinhardii 48 17 21

I. GENERAL
1.5 BIOFUEL PRODUCTION 17
TABLE 1.7 Chemical Compositions of Macroalgae and Microalgae (wt% Dry Weight) [14]—cont’d
Algae Protein Carbohydrate Lipid
Chlorella vulgaris 51–58 12–17 14–22

Chlorella pyrenoidosa 57 26 2
Spirogyra sp. 6–20 33–64 11–21
Dunaliella bioculata 49 4 8
Dunaliella salina 57 32 6
Euglena gracilis 39–61 14–18 14–20
Prymnesium parvum 28–45 25–33 22–39

Tetraselmis maculata 52 15 3
Porphyridium cruentum 28–39 40–57 9–14
Spirulina platensis 46–63 8–14 4–9
Spirulina maxima 60–71 13–16 6–7
Synechoccus sp. 63 15 11

Anabaena cylindrica 43–56 25–30 4–7

1.5 BIOFUEL PRODUCTION

Generally, the following technologies can be used to process or produce a biofuel:


(a) physical technologies, (b) chemical technologies, (c) thermochemical technologies, and
(d) biochemical technologies. Biological technologies such as sequestration and
biosequestration can also be used to produce biofuels, for example, by using algae, bacteria,
and cyanobacteria to capture and store the atmospheric GHG (CO2) via biological processes
[84]. Sometimes, multiple technologies combining the above technologies are also used.

1.5.1 Physical Technologies


For some special biofuels, some physical technologies can be used, for example, mechan-
ical extraction can be used to extract biooil from oil crops; distillation can be used to separate
biooil from impurities; and briquetting can be used to densify a solid biofuel, etc.

1.5.1.1 Physical Extraction


Physical extraction is also called mechanical extraction in some publications [35]. Oilseed
crops with high oil content can be physically crushed, and the oils can be extracted and
converted to esters which then can be directly used to replace diesel or as heating oil. There
is a wide range of oilseed crops that can be used for biodiesel production, and the most com-
monly used crop is rapeseed. Other feedstocks used include cotton, groundnuts, palm, sun-
flower, soya bean, and recycled frying oil [85].

I. GENERAL
18 1. BIOFUELS: INTRODUCTION

Mechanical pressure using screw press (expeller) or extractor can be used to recover crude
vegetable oils from the oilseeds. Screw press can be applied in two ways: prepressing and full
pressing. In the prepressing process, only part of the oil is recovered and the partially de-oiled
meal (generally contains 18%–20% of oil) can be further extracted using solvents like hexane,
ethanol, 2-propanol, chloroform, or toluene [86]. Combined prepressing and solvent extrac-
tion is commonly applied for oilseeds with high oil content (30%–40%). Full pressing requires
95 MPa to squeeze out as much oil as possible, preferably up to 3%–5% residual fat for animal
materials. Full pressing can also be carried out in a prepress and a final press stage [35].

1.5.1.2 Distillation
Distillation is the process of separating the components or substances from a liquid
mixture by selective boiling and condensation. Distillation may result in essentially complete
separation (nearly pure components), or it may be a partial separation that increases the con-
centration of selected components of the mixture. In either case, the process exploits differ-
ences in the volatilities of the mixture components.
Distillation is an important method for extracting essential oils from oil mixtures. Initially,
oil plants are crushed to release the oils, and they are then steam distilled. During this process,
the essential oils vaporize and rise up with the steam. The oil vapors are captured, and are
allowed to condense back into biooil liquids [85]. To achieve a better distillation, advanced
technologies can be used, for example, vacuum distillation [87], molecular distillation
[88], etc.

1.5.1.3 Densification
Solid biofuels like agricultural residues, forestry residues, and other waste materials are
often difficult to be directly used because the fuel characteristics are generally troublesome,
for example, uneven particle size, low bulk density, low energy content, etc. These disad-
vantages can be efficiently overcome by means of densification of the solids into compact
regular shapes. For some coarse or fluffy materials (e.g., tree branches, cereal straws),
chopping, grinding, and pulverizing are efficient approaches to produce high-quality fuels
[85, 89]. Sometimes, briquetting or pelletization can further be used to improve the biofuel
properties [55, 90].

1.5.2 Chemical Technologies


Some chemical technologies can also be used to produce or obtain biofuels, which mainly
include solvent extraction, supercritical (fluid) extraction, supercritical water conversion, etc.

1.5.2.1 Solvent Extraction


Extraction refers to a process in which the desired substance is selectively removed from
the raw materials by allowing the desired substance to dissolve into the solvent, and subse-
quently recovering the substance from the solvent [35]. To remove the particular substance
from biomass, extraction and separation are both essential. Typically, biomass (wood, wheat
straw, aromatic grasses, etc.) contains high volume of macromolecular compounds (polysac-
charide, cellulose, hemicellulose, lignin, etc.) which are called primary metabolite. The other

I. GENERAL
1.5 BIOFUEL PRODUCTION 19
low-volume and high-value chemical molecules like terpenoids, waxes, resins, sterols, and
alkaloids are known as secondary metabolites or extractive biomass. In the biorefinery pro-
cess these chemicals are initially extracted from biomass by using solvent extraction [35].
Solvent extraction generally involves different unit operations: (a) extract the oil from the
oilseeds using hexane as a solvent, (b) evaporate the solvent, (c) distil the oil-hexane mixture
(also called miscella), and (d) toast the de-oiled meal. In some special cases, the other solvents
can be used, for example, acetone, ethanol, isopropanol, halogenated solvents (mostly
dichloromethane), etc. [35].

1.5.2.2 Supercritical Extraction


Supercritical extraction is also called supercritical fluid extraction. A supercritical fluid is
defined as a substance that is under temperature and pressure conditions which are above its
vapor liquid critical point (for water it is 644 K and 22 MPa; for CO2 it is 304 K and 7.4 MPa).
Under these conditions, the supercritical fluid has some important advantages, for example,
its physical properties (e.g., density, solubility, diffusion coefficient, dielectric constant, etc.)
can be easily manipulated through slight variations in the temperature and pressure, making
the supercritical fluid an excellent medium for extraction and reaction [50, 91].
In biorefinery processes, the chemical molecules including terpenoids, waxes, resins, ste-
rols, polycosanol, and alkaloids (known as secondary metabolites or extractive biomass) can
be extracted from biomass using supercritical extraction [35]. During the process, aromatic
woods (cedar wood, sandal wood, pine wood, etc.) are extracted to isolate extractives, the
extracted lignocellulosic biomass is further used for hydrolysis and fermentation for produc-
ing biofuels. Supercritical extraction is usually performed using CO2 [35, 50].

1.5.2.3 Supercritical Water Conversion


Supercritical water conversion of biomass to biofuels represents an alternative path to
acid hydrolysis and enzymatic hydrolysis. Supercritical water can quickly convert cellulose
to sugar and biomass into a mixture of oils, organic acids, alcohol, and methane. At super-
critical (e.g., temperature of 300–644 K, pressure of 200–250 bar) or near critical states
(523–573 K), acid (H+) and base components (OH) of water are separated and dissolved
in the biomass. The dissolved supercritical water rapidly breaks the bonds of cellulose
and hemicellulose to produce small sugar molecules like glucose, xylose, and oligosaccha-
ride [35]. These properties make supercritical water a very promising reaction medium
without using any catalyst for the conversion of biomass to value-added biofuels. Super-
critical water also breaks the cellulosic bonds and forms gaseous products. The typical
overall reaction is [35].

2C6 H12 O6 + 7H2 O ! CO2 + 2CH4 + CO + 5H2 (1.1)

1.5.3 Thermochemical Technologies


As compared with the other technologies including physical technologies, chemical tech-
nologies, and biochemical technologies, thermochemical technologies are generally more
versatile and can essentially convert all the organic components of the feedstock into a range
of products which are mainly biofuels [51]. Generally, thermochemical (or thermo-chemical)

I. GENERAL
20 1. BIOFUELS: INTRODUCTION

technologies mainly include combustion, gasification, pyrolysis, liquidation, etc. In some


publications, some other methods are also included, for example, incineration [92],
torrefaction [60], hydrogenation [93, 94], etc.

1.5.3.1 Combustion
Combustion is a thermochemical process where fuel is burnt in oxygen-excess atmosphere (air
or oxygen), and the chemical energy stored in the fuel is released to produce heat which can be
used for cooking, space heating, and electricity generation. The overall reactions are as follows:
Biofuel + O2 ! CO2 + H2 O + heat (1.2)
Biofuel + Air ! CO2 + H2 O + N2 + heat (1.3)
If the biofuel contains ash, S, or N, ash, SOx, or NOx would also be included in the above
equations [95, 96].
For theoretical combustion, the C and H contents in the fuel would be converted to CO2 and
H2O, respectively, and the heat produced can result in a high temperature of 2000°C (theoret-
ical). Practically, the combustion would be incomplete thereby leading to the production of CO,
and the heat would be diffused (to the environment) thereby leading to relatively low temper-
ature of 800–1000°C [85]. To help a better combustion, excess air or oxygen is usually used [52].
Theoretically, it is possible to (completely) combust any type of biomass or biofuel, but
practically, combustion is feasible only when the fuel has moisture content of <50% [85,
93]. To resolve this issue, a preheating step is usually used.
Although combustion is an efficient and mature technology to convert biofuel and gener-
ate heat, it is actually not a technology to produce biofuel because no biofuels would be pro-
duced after the complete combustion of the original feedstock (mainly CO2 and H2O are left).

1.5.3.2 Gasification
Gasification is also a thermochemical process where the reactions between fuel and gasi-
fication agent take place and syngas (also known as producer gas, product gas, synthetic gas,
or synthesis gas) is produced. The syngas is mainly composed of CO, H2, N2, CO2, and some
hydrocarbons (CH4, C2H4, C2H6, etc.). Very small amounts of H2S, NH3, and tars may also be
produced [97].
Based on the gasification agents used, biomass gasification processes can be divided into
air gasification (using air [98]), oxygen gasification (using oxygen [99]), steam gasification
(using steam [100]), carbon dioxide gasification (using carbon dioxide [101]), supercritical
water gasification (using supercritical water [102]), etc. Generally, oxygen gasification, steam
gasification, carbon dioxide gasification, and supercritical water gasification result in higher
heating values (HHVs) of syngas than air gasification, however, air gasification is the most
widely studied and applied because the gasification agent (air) is cheap, the reaction process
is easy, the reactor structure is simple, etc.
For air gasification of a biofuel, the following reactions take place during the gasification
process [92, 103, 104]:
Drying or moisture release:
Wet biofuel ! dry biofuel + H2 O (1.4)
Devolatilization:

I. GENERAL
1.5 BIOFUEL PRODUCTION 21

Dry biofuel ! CO,CO2 , CH4 , C2 H4 , H2 O + Carbon and Primary tar CHx Oy (1.5)
Cracking and reforming:

Primary tar ! CO,CO2 , CH4 , C2 H4 , H2 + Secondary tar (1.6)

Homogeneous gas phase reactions:

Secondary tars ! C, CO,H2 (1.7)

Dry reforming reaction : CH4 + CO2 ! 2CO + 2H2 ΔH ¼ + 247kJ=mol (1.8)

Steam reforming methanization : CH4 + H2 O ! CO + 3H2 ΔH ¼ + 206kJ=mol (1.9)

Water  gas  shift reaction : CO + H2 O ! CO2 + H2 ΔH ¼ 40:9kJ=mol (1.10)

Combustion : CO + 0:5O2 ! CO2 ΔH ¼ 283kJ=mol (1.11)

Combustion : H2 + 0:5O2 ! H2 O ΔH ¼ 242kJ=mol (1.12)

Combustion : CH4 + 0:5O2 ! CO + 2H2 ΔH ¼ 110kJ=mol (1.13)

Heterogeneous reactions:
Combustion : C + O2 ! CO2 ΔH ¼ 393:5kJ=mol (1.14)
Combustion : C + 0:5O2 ! CO ΔH ¼ 123:1kJ=mol (1.15)
Boudouard equilibrium reaction : C + CO2 ! 2CO ΔH ¼ + 159:9kJ=mol (1.16)

Water  gas reaction steam reforming : C + H2 O ! CO + H2 ΔH ¼ + 118:5kJ=mol (1.17)

Methane production reaction : C + 2H2 ! CH4 ΔH ¼ 87:5kJ=mol (1.18)

For biofuel gasification with steam, carbon dioxide, or supercritical water, the overall gas-
ification reaction is generally endothermic, external heating is therefore required during the
whole gasification process. However, for biofuel gasification with air or oxygen, the overall
gasification may be endothermic or exothermic. These can be controlled or changed by vary-
ing the air or oxygen content which is represented by equivalence ratio (ER, defined as the
ratio of air or oxygen used for gasification by the air or oxygen required for stoichiometric
(theoretically complete) combustion). Generally, a specific ER corresponds to a specific gas-
ification temperature if no external heat is provided. If higher gasification temperature is re-
quired or designed, external heat should be input or a higher ER is needed [105].
A unit in which the gasification reactions take place is known as a gasifier. A gasifier is a
very important factor that affects the gasification processes, reactions, and products. Gener-
ally, gasifiers can be classified into three broad groups: fixed bed gasifiers (or moving bed
gasifiers), fluidized bed gasifiers, and entrained flow gasifiers [97]. The fixed bed gasifiers
mainly include updraft, downdraft, horizontal-draft, etc. The fluidized bed gasifiers mainly
include bubbling fluidized bed, circulated fluidized bed, double circulated fluidized bed, etc.

I. GENERAL
22 1. BIOFUELS: INTRODUCTION

A review of the 50 gasifier manufacturers in Europe, United States, and Canada showed that
75% of the designs were downdraft fixed beds, 2.5% were updraft fixed beds, 20% were flu-
idized beds, and 2.5% were the other designs [97, 106]. Table 1.8 presents the main character-
istics of different gasifiers, with which, specific gasifiers can be selected and designed.

1.5.3.3 Pyrolysis
Pyrolysis is typically defined as a thermochemical decomposition of biomass feedstock at
medium temperatures (300–800°C) to high temperatures (800–1300°C) in the inert atmo-
sphere [109]. Some similar definitions are reported in other publications [7, 35, 52, 61, 85,
93, 110].
The overall reaction of biomass feedstock is:
Biofuel + heat ! liquid + syngas + solid (1.19)
The products of liquid, syngas, and solid are actually all valuable fuels and are defined as
biooil (or bio-oil), biosyngas (or bio-syngas), and biochar (or bio-char), respectively.
The chemical reactions during the pyrolysis process mainly include [108] the following:
C + 2H2 O ! 2H2 + CO2 ΔH ¼ + 75kJ=mol (1.20)
C + H2 O ! H2 + CO ΔH ¼ + 131kJ=mol (1.21)

TABLE 1.8 Main Characteristics of Different Gasifiers [97, 107, 108]


Gasifier Characteristics
Fixed bed (a) Small capacities (0.01–10 MW)
(b) Can handle large and coarse particles
(c) Low product gas temperature (450–650°C)
(d) High particulate content in gas product stream
(e) High gasifying agent consumption
(f ) Ash is removed as slag or dry
(g) May result in high tar content (0.01–150 g/Nm3)
Fluidized bed (a) Medium-size capacities (1–100 MW)
(b) Uniform temperature distribution
(c) Better gas–solid contact
(d) High operating temperature (1000–1200°C)
(e) Low particulate content in the gas stream
(f ) Suitable for feedstocks with low ash fusion temperature
(g) Ash is removed as slag or dry
Entrained flow (a) Large capacities (60–1000 MW)
(b) Needs finely divided feed material (<0.1–0.4 mm)
(c) Very high operating temperatures (>1200°C)
(d) Not suitable for high ash content feedstocks
(e) Very high oxygen demand
(f ) Short residence time
(g) Ash is removed as slag
(h) May result in low tar content (negligible)

I. GENERAL
1.5 BIOFUEL PRODUCTION 23
CH4 + 2H2 O ! 4H2 + CO2 ΔH ¼ + 165kJ=mol (1.22)
CH4 + H2 O ! CO + 3H2 ΔH ¼ + 206kJ=mol (1.23)
As compared with the other thermochemical technologies including combustion, gasifica-
tion, and liquidation, pyrolysis generally has the following advantages: (a) the main product
is biooil and the yield may be as high as 75%, (b) the biooil may have high content of carbon,
(c) the biooil may have low contents of nitrogen and sulfur, (d) the HHV of biooil may be very
high, for example, 42 MJ/kg, which is comparable with those of fossil fuels, (e) the residence
time is generally short which decreases the operational cost, (f ) desired product (biooil,
biosyngas, or biochar) could be produced by adjusting the operational parameters,
(g) biooil could be easily stored or transported, and (h) biomass feedstock may not need to
be processed [111].
According to the operating parameters such as heating rate, pyrolysis temperature, and
residence time, conventional electrical pyrolysis can be generally classified into three groups:
(a) slow pyrolysis, (b) fast pyrolysis, and (c) flash pyrolysis. In some publications, catalytic
pyrolysis, microwave pyrolysis, vacuum pyrolysis, and hydropyrolysis are also included
[35, 51, 111]. For slow pyrolysis, the heating rate, pyrolysis temperature, and residence time
are mainly <1°C/s, 300–700°C, and >450 s, respectively. For fast pyrolysis, they are mainly
10–300°C/s, 550–1250°C, and 0.5–20 s, respectively. For flash pyrolysis, they are >1000°C/s,
800–1300°C, and <0.5 s, respectively [109, 112]. These different conditions generally result in
different pyrolysis results. Table 1.9 presents the pyrolysis results obtained from different py-
rolysis technologies. It is observed that flash pyrolysis generally favors biooil production,
followed by fast pyrolysis and slow pyrolysis.
Recently, a new pyrolysis technology, microwave-assisted pyrolysis (MAP), was devel-
oped and widely used [114–116], and it has drawn serious attention due to its advantages
over the conventional electrical pyrolysis which are illustrated in Table 1.10. These advan-
tages are due to the different heat and mass transfer mechanisms. For the conventional elec-
trical pyrolysis, heat is transferred from high-temperature gas to the fuel particle surface
through convection mechanism and is then further transferred from the outside surface to

TABLE 1.9 Pyrolysis Results Obtained From Different Pyrolysis Technologies [113]
Pyrolysis Operating Conditions Results (wt%)
Slow pyrolysis Temperature: 300–700°C Bio-oil: 30
Vapor residence time: 10–100 min Biochar: 35
Heating rate: 0.1–1°C/s Gases: 35
Feedstock size: 5–50 mm
Fast pyrolysis Temperature: 400–800°C Bio-oil: 50
Vapor residence time: 0.5–5 s Biochar: 20
Heating rate: 10–200°C/s Gases: 30
Feedstock size: <3 mm
Flash pyrolysis Temperature: 800–1000°C Bio-oil: 75
Vapor residence time: <0.5 s Biochar: 12
Heating rate: >1000°C/s Gases: 13
Feedstock size: <0.2 mm

I. GENERAL
24 1. BIOFUELS: INTRODUCTION

TABLE 1.10 Comparison Between Microwave-Assisted Pyrolysis and Conventional


Electrical Pyrolysis [109, 117]
Microwave-Assisted Pyrolysis Conventional Electrical Pyrolysis
Conversion of energy Transfer of energy
Noncontact heating Contact heating

Hot spot No hot spot


Selective Nonselective
Lower thermal inertia and faster response Higher thermal inertia and slower response
Lower energy consumption Higher energy consumption
Rapid heating Slow heating
Shorter reaction times Longer reaction times

Volumetric heating Superficial heating


Higher level of control Lower level of control
Improved product yields Lower product yields

the inside core through conduction mechanism. A temperature gradient from outside to in-
side of the feedstock particle is formed, and the released volatile diffuses from the inside
core to the outside surface through a higher temperature region. For the MAP, microwave
penetrates the feedstock particle and microwave energy is transformed into thermal energy
which constantly accumulates inside the biomass particle and is then transferred outwards.
A temperature gradient from inside to outside of particle is formed, and the released vol-
atile diffuses from the inside core to the outside surface through a lower temperature
region [109].

1.5.3.4 Liquefaction
Liquefaction is a thermochemical process where biomass undergoes complicated chemical
reactions in a solvent medium to form mainly liquid products (biooil or bio-oil). Hydrother-
mal liquefaction is a process in which water is used as reaction medium, and the process is
carried out in sub-/supercritical water (200–400°C) under sufficient pressure to liquefy bio-
mass for biooil production [80].
Alkali salts such as sodium carbonate and potassium carbonate can act as catalyst for
hydrolysis of macromolecules like cellulose and hemicellulose into smaller fragments.
The heavy oil obtained from the liquefaction process is a viscous tarry lump, which
may sometimes cause handling troubles. In this case, some organic solvents (e.g., propanol,
butanol, acetone, methyl ethyl ketone, ethyl acetate, etc.) are added to the reaction system.
Generally, catalytic aqueous liquefaction may result in higher biooil yield than the
noncatalytic aqueous liquefaction [35]. Compared with biooil obtained from pyrolysis
method, the biooil yield from liquefaction method is generally much lower and the biooil
is more viscous [35].

I. GENERAL
1.5 BIOFUEL PRODUCTION 25
Biooils obtained from liquefaction process generally contain high contents of volatile or-
ganic acids, alcohols, aldehydes, ethers, esters, ketones, and nonvolatile components. These
oil components could be catalytically upgraded to yield an organic distillate product which is
rich in hydrocarbons and useful chemicals [35].
On the other hand, also when compared with pyrolysis, the feedstock feeding
system for liquefaction is more complex and more expensive, rendering a low interest
in liquefaction [85, 93].

1.5.4 Biochemical Technologies


The biochemical technologies for biomass conversion mainly include digestion, fermenta-
tion, and hydrolysis. In some publications, interesterification [118] and transesterification [94]
are also included. In practical applications, the above technologies are usually combined for
the production of a biofuel [119].

1.5.4.1 Digestion
Biomass digestion utilizes bacteria to facilitate the decomposition of biomass feedstock and
this may occur under anaerobic (anaerobic digestion) or aerobic (aerobic digestion) condi-
tions. Compared with aerobic digestion, anaerobic digestion has some advantages:
(a) higher organic removal efficiency, (b) higher organic loading rate, (c) lower nutrient re-
quirement, and (d) lower energy requirement. On the other hand, aerobic digestion may
not facilitate the conversion of biomass feedstock to any high-value products. Anaerobic
digestion is therefore more frequently considered or utilized [92].
Anaerobic digestion is a biochemical conversion of biomass in which numerous species of
bacteria participate in the decomposition of organic matter in the absence of oxygen to pro-
duce a biogas which contains approximately 55%–75% methane and 25%–45% carbon dioxide
depending on the water content. Trace gases such as hydrogen sulfide may also be
produced [85].
During an anaerobic digestion process, specialized microorganisms break down the com-
plex organic matter (carbohydrates, proteins, and fats) into smaller molecules (sugars, amino
acids, and fatty acids) that are soluble in water. Methane and carbon dioxide are the primary
and also the final products of this process (known as biogas). The overall reaction for anaer-
obic digestion is:
C6 H12 O6 ! 3CO2 + 3CH4 (1.24)
The anaerobic digestion process can be generally divided into four steps [120]:
(a) Hydrolysis: the large organic polymers (proteins, fats, carbohydrates, etc.) are broken
down into smaller molecules (amino acids, fatty acids, simple sugars, etc.).
(b) Acidogenesis: acidogenic microorganisms (acidogenic bacteria) further break down the
organic matter to produce an acidic environment.
(c) Acetogenesis: acetogens catabolize many of the products created in acidogenesis into
acetic acid, CO2, and H2.

I. GENERAL
26 1. BIOFUELS: INTRODUCTION

(d) Methanogenesis: methanogens create methane from the final products of acetogenesis, as
well as from some of the intermediate products from hydrolysis and acidogenesis mainly
through two pathways:

CO2 + 4H2 ! CH4 + 2H2 O (1.25)


CH3 COOH ! CH4 + CO2 (1.26)
Anaerobic digestion is a versatile process that can be applied to a wide variety of biomass
feedstocks, including municipal solid waste, industrial waste, livestock manure, food
processing waste, as well as domestic and industrial sewages which generally have high
moisture contents (80–90 wt% of biomass feedstock) [85].

1.5.4.2 Hydrolysis
During the process of converting lignocellulosic biomass to bioethanol, lignocellulosic
biomass is initially hydrolyzed to fermentable sugars, and then fermentation is followed to
produce alcohol (bioethanol). For this whole process, hydrolysis (or saccharification) of lig-
nocellulosic material into fermentable sugars is a crucial stage, which determines the overall
process efficiency [121].
There are several methods that can be used to derive sugars from lignocellulosic biomass,
for example, physical, chemical, and enzymatic. Among these methods, chemical hydrolysis
and enzymatic hydrolysis are the most successful ones.
Chemical hydrolysis can be performed using concentrated mineral acids such as H2SO4
and HCl. It can also be performed using less hazardous and more tractable cellulose solvents
such as ionic liquids. The acid hydrolysis approach is comparatively cheap, however, its
application is limited due to the low yields and unfavorable environmental issues involved
with the use of strong acids [61, 121].
Enzymatic hydrolysis is usually carried out by highly specific cellulase enzymes, and the
products are usually reducing sugars including glucose. Compared with chemical hydrolysis
(performed using strong acids), enzymatic hydrolysis is mainly conducted under milder
conditions (e.g., optimum pH of 4.8 and temperature of 45–50°C for cellulase enzyme). Also,
enzymatic hydrolysis has some advantages: (a) high product yield, (b) less or no by-products,
(c) high selectivity, and (d) no corrosion problem [61, 121]. However, enzymatic hydrolysis
may take a much longer time (several days vs a few minutes), and the final product may
inhibit the enzyme [121]. Currently, a major bottleneck in enzymatic hydrolysis is the high
cost of the enzymes [61, 121].

1.5.4.3 Fermentation
During the process of converting lignocellulosic biomass to bioethanol, sugars are first de-
rived from hydrolysis of the lignocellulosic biomass, and then further treated by an important
process to generate alcohol (bioethanol)—fermentation.
The term “fermentation” is derived from the Latin verb “fervere” (means to boil), which
describes the action of yeast on the extracts of fruit or malted grain: boiling appears when
carbon dioxide bubbles are produced by the anaerobic catabolism of sugar in the extract [122].

I. GENERAL
1.6 SOLID BIOFUELS 27
Broadly, fermentation is a metabolic process that consumes sugar in the absence of oxygen.
It generally occurs in yeast, bacteria, or oxygen-starved muscle cells, and the products are
alcohol, gases, or organic acids [123].
Alcohol and gas are mainly produced through ethanol fermentation [124]:
C6 H12 O6 ! 2C2 H5 OH + 2CO2 (1.27)
Organic acids are mainly produced through homolactic fermentation and heterolactic
fermentation [123]:
Homolactic : C6 H12 O6 ! 2CH3 CHOHCOOH (1.28)
Heterolactic : C6 H12 O6 ! CH3 CHOHCOOH + C2 H5 OH + CO2 (1.29)
Once produced, ethanol generally needs to be distilled and dehydrated to produce a high-
octane, water-free alcohol [85].
Currently, fermentation is commercially used on a large scale in various countries to pro-
duce bioethanol from biomass feedstocks. The best-known biomass feedstocks are sugar
crops and starch crops. Lignocellulosic biomass feedstocks such as cereals and woods can also
be processed [85].

1.6 SOLID BIOFUELS

1.6.1 Solid Biomass


Solid biomass mainly refers to the raw solid biomass feedstock that is used for biofuel pro-
duction, for example, agricultural residues, forest residues, energy crops, and solid wastes.
Actually, the raw solid biomass feedstock is also a biofuel. As stated in Section 1.4, solid
biomass mainly includes oil crops, lignocellulosic biomass, solid wastes, etc. Theoretically,
all these materials can be converted to biofuels and can also be used as biofuels directly.

1.6.2 Biochar
Compared with fossil fuels, the raw feedstocks, for example, woods (firewood, wood
chips, or wood pellets) generally have higher combustion emissions and lower energy con-
tents (heating values). Also, the fire temperature from wood combustion would be very low
(e.g., below 850°C), making it hard to melt metals [1]. To overcome these drawbacks, biochar
was developed and used.
Biochar is also known as charcoal, which is a porous, carbon-enriched, grayish black
solid [1]. It is typically produced by heating biomass feedstock in kiln or retort (pyrolysis)
at the temperatures of around 400°C (generally between 300°C and 900°C) in the absence
of air [1,125]. It can also be produced from torrefaction, gasification, hydrothermal
carbonization, etc.
Compared with the raw feedstocks, biochars have some advantages. Table 1.11 presents
the main characteristics including H/C, O/C, and HHV (higher heating value) of some
raw feedstocks and the produced biochars. It is observed that the H/C, O/C, and HHV of

I. GENERAL
28 1. BIOFUELS: INTRODUCTION

TABLE 1.11 Characteristics of Different Biochars [125]


Feedstock Biochar
Feedstock H/C O/C HHV (MJ/kg) H/C O/C HHV (MJ/kg)
White pine 1.40 0.62 18.06 0.43 0.08 31.41
Fruit cuttings 1.61 0.77 17.27 0.12 0.09 29.02
Switchgrass 1.56 0.77 19.50 0.70 0.24 26.60
Barley straw 1.45 0.66 17.34 0.40 0.11 26.02

Safflower seeds 1.68 0.62 24.80 0.43 0.21 30.27


Eucalyptus sawdust 1.47 0.71 16.69 0.81 0.27 26.19
Barley straw 1.45 0.66 17.34 0.86 0.21 27.49
Corn stover 1.62 0.73 16.20 0.94 0.18 27.76
Spruce 1.49 0.65 19.94 1.24 0.49 22.50
Birch 1.56 0.68 20.42 1.24 0.49 22.50

Maize silage 1.58 0.55 22.30 1.13 0.09 35.70


Coconut fiber 1.41 0.71 18.40 0.66 0.15 30.60
Eucalyptus leaves 1.59 0.72 18.90 1.01 0.22 29.40
Corn stalk 1.58 0.65 17.51 0.94 0.17 29.21
Wood 1.65 0.68 17.93 0.90 0.22 28.38

the raw feedstocks are 1.40–1.68, 0.55–0.77, 16.20–24.80 MJ/kg, respectively. The H/C, O/C,
and HHV of the biochars are 0.12–1.24, 0.08–0.49, 22.50–35.70 MJ/kg, respectively. Generally,
biochars have much higher HHVs and much lower H/C and O/C ratios than the raw feed-
stocks. These are mainly caused by the increases in the carbon content. For instance, the fixed
carbon content has increased from 16.39 to 81.03 wt% (4.94 times) when the white pine was
pyrolyzed to biochar [125].
Compared with the raw feedstocks, the obtained high-quality biochars have much
more usages including: (a) solid fuel in boilers, (b) feedstock in further gasification process,
(c) carbon sequestration, (d) production of activated carbon, (e) making carbon-nanotubes,
(f ) making briquettes, etc. [93, 125].

1.7 LIQUID BIOFUELS

Liquid biofuels are receiving more and more attention and interest due to the facts that:
(a) they are liquid (and easy for storage), (b) they are highly combustible, (c) they are
nonexplosive, (d) they have high energy to mass ratios, (e) they are stable for long-term stor-
age, (f ) they can be transported via pipelines, and (g) they are inexpensive [5]. Liquid biofuels
mainly include biomethanol, bioethanol, biobutanol, biopropanol, biodiesel, jet fuel, etc.
Among these liquid biofuels, bioethanol, biodiesel, and biooil are the most important ones.

I. GENERAL
1.7 LIQUID BIOFUELS 29

1.7.1 Bioethanol
Bioethanol is also known as ethylic alcohol or ethanol. Its molecular formula is
CH3dCH2dOH which is the same organic compound that is used in alcoholic beverages.
Nowadays, bioethanol is the most commonly used liquid biofuel in the world.
Bioethanol is mainly produced by microbial fermentation of sugar or starch from various
feedstocks, including sugarcane, corn, grains, agricultural wastes, forestry wastes, municipal
wastes, livestock manures, etc. Because the raw feedstock usually has lignin content, hydro-
lysis process (acid or enzymatic) is usually used for delignification before the fermentation.
With the help of microorganisms (e.g., yeast), the hydrolyzed sugars (e.g., glucose) are
converted into ethanol (bioethanol) during the fermentation process. Distillation and dehy-
dration are then used to produce a high-octane, water-free alcohol, which is also known as
hydrated ethanol or anhydrous ethanol [85, 126].
The produced bioethanol can be directly used as gasoline substitute in engines. Table 1.12
illustrates the physicochemical properties of gasoline and ethanol (bioethanol). It is observed
that ethanol has a lower LHV (21.1 MJ/L) than gasoline (30–33 MJ/L), therefore more ethanol
is required to obtain the same output. However, the higher octane number of ethanol allows a
higher engine compression ratio to be used, which leads to improved thermal efficiency and
increased power, thereby reducing somewhat the difference in fuel consumption [4]. It is also

TABLE 1.12 Physicochemical Properties of Gasoline and Ethanol [127]


Property Gasoline Ethanol
Formula C4 to C12 C2H5OH
Molecular weight 100–105 46.07

Density at 15°C (kg/L) 0.69–0.79 0.79


Specific gravity (relative density) at 15°C 91 106–110
Freezing point (°C) 40 114
Boiling point (°C) 27–225 78
Vapor pressure at 38°C (kPa) 48–103 15.9
Specific heat (kJ/kg K) 2.0 2.4

Viscosity at 20°C (mPa s) 0.37–0.44 1.19


Lower heating value (MJ/L) 30–33 21.1
Flash point (°C) 43 13
Auto-ignition temperature (°C) 257 423
Lower flammability limit (vol%) 1.4 4.3
Higher flammability limit (vol%) 7.6 19.0

Stoichiometric air-fuel ratio 14.7 9.0


Research octane number 88–100 108.6
Motor octane number 80–90 89.7

I. GENERAL
30 1. BIOFUELS: INTRODUCTION

observed that there are significant differences among the other properties of the two fuels, for
example, viscosity, specific gravity, freezing point, boiling point, flash point, vapor pressure,
flammability limit, auto-ignition temperature, stoichiometric air-fuel ratio, etc.
To improve the fuel quality such as volatility, octane number, cold start, hot operation, and
fuel consumption, bioethanol is usually blended with gasoline, for example, bends of E5 (5%
ethanol and 95% gasoline) to E25 (25% ethanol and 75% gasoline) require no alterations on the
engine equipment or settings [35].
Also, bioethanol can be used in tractors, planes, and boats [37]. It can be used as an octane
enhancer to replace tetraethyl lead [5]. It can be used as valuable chemicals for ethylene and
ethylene glycol production. It can also be used as additives to increase the fuel oxygen per-
centage, which reduces the CO and aromatic emissions [8].

1.7.2 Biodiesel
As the two most important liquid biofuels used in transportation sectors, bioethanol is a
substitute for gasoline whereas biodiesel is a substitute for diesel.
Biodiesel is a nearly colorless (sometimes yellowish) liquid mixture of long-chain fatty acid
methyl esters (FAME) mainly derived from vegetable oil, animal fats, algal lipids, or waste
grease through transesterification (alcoholysis) in the presence of alcohol and alkaline
catalysts [1, 5, 7, 48]. Typically, biodiesel encompasses alkyl fatty acid esters of short-chain
alcohols (primarily methanol or ethanol) [48], and its chemical structure is a chain of 12–22
carbon atoms (average is 16) and 0–2 double bonds [5, 7]. The chemical structure of biodiesel
is different from that of regular diesel. Table 1.13 illustrates the physical and chemical prop-
erties of diesel and biodiesel. It is shown that petroleum diesel contains only hydrocarbons
(hydrogen and carbon, no oxygen) whereas biodiesel contains carbon, hydrogen, and oxygen.
Also, there are differences in the chemical compounds, for example, petroleum diesel is
mainly composed of n-aliphatics (67.4%) and aromatics (20.1%) whereas biodiesel is mainly
composed of olephenics (84.7%) and n-aliphatics (15.2%). Although petroleum diesel and
biodiesel have similar densities, the above composition differences make the two fuels dis-
tinct in the other properties, which include both advantages and disadvantages. The advan-
tages of biodiesel mainly include its: (a) higher flash point which makes it safer for storage,
handling, use, and transportation, and (b) higher cetane number which means that it has a
better ignition quality. The disadvantages of biodiesel mainly include its: (a) lower calorific
value which means more biodiesel will be consumed to supply the same output, (b) higher
pour point which means it tends to gel up (solidify) in cold weather, and (c) higher viscosity
which has a negative effect on fuel spray atomization. Some other disadvantages of biodiesel
include the facts that the acids will harm the engine parts, and the water content will lower the
energy quality.
On the whole, biodiesel is a better fuel than petroleum diesel, because: (a) the feedstocks
are renewable, (b) it can be combusted more completely, (c) it does not yield explosive vapors,
(d) it is safer due to its higher flash point, (e) it is safer for the environment due to its lower
exhaust emission profile, (f ) it may be less toxic due to its lower or negligible sulfur content,
and (g) it is biodegradable [5, 48, 129]. Therefore, it can be directly used (without any mod-
ifications) in home heating as well as in car engines, traction motors, railway locomotives,

I. GENERAL
1.7 LIQUID BIOFUELS 31
TABLE 1.13 Physical and Chemical Properties of Diesel and Biodiesel [18, 128]
Property Diesel Biodiesel
Carbon (%) 86.4 79.6

Hydrogen (%) 13.6 10.5


Oxygen (%) – 8.6
Nitrogen (%) – 1.3
C/H 6.5 7.6
n-Aliphatics (%) 67.4 15.2
Olephenics (%) 3.4 84.7

Aromatics (%) 20.1 –


Naphtens (%) 9.1 –
3
Density at288 K (kg/m ) 816–840 820–897
Calorific value (MJ/kg) 42–46 32–43
Flash point (K) 323–371 408–469

Pour point (K) 238–258 260–289


2
Kinematic viscosity at 300 K (mm /s) 2.5–5.7 4–26
Cetane number 45–55 46–66

construction machineries, mining equipment, military carriers, etc., especially in warm cli-
mates [1, 37, 130]. To balance some of the disadvantages of biodiesel, biodiesel is usually
blended with regular crude oil, petroleum diesel, alcohol (methanol), or microemulsions,
and any proportion can be used. Currently, the most common biodiesel blends are B2
(2% biodiesel and 98% regular diesel), B5 (5% biodiesel and 95% regular diesel), and B20
(20% biodiesel and 80% regular diesel) [37].

1.7.3 Biooil
Biooil is a dark brown and free-flowing organic liquid. Its synonyms mainly include py-
rolysis oil, pyrolysis liquid, bio-crude oil, wood oil, wood distillate, wood liquid, liquid
smoke, pyroligneous acid, pyrolytic tar, and so on [15].
Pyrolysis biooil is a complex mixture of >300 components which are mainly derived from
depolymerization or fragmentation of biomass. It generally has different compositions and
properties as compared with diesel oils and petroleum oils. Table 1.14 presents the main
physical and chemical properties of the conventional electrical pyrolysis biooil, MAP biooil,
diesel oil, and petroleum oil. It is observed that the pyrolysis biooils (both conventional
electrical pyrolysis biooil and MAP biooil) generally have much higher moisture contents
(4.5–43.0 wt%) than diesel oil (0 wt%) and petroleum oil (0.1 wt%). These high moisture con-
tents are mainly from two sources: (a) the water in the raw feedstock and (b) the water

I. GENERAL
32 1. BIOFUELS: INTRODUCTION

TABLE 1.14 Physical and Chemical Properties of Different Oils [15]


Pyrolysis Biooil
Property Microwave Conventional Diesel Oil Petroleum Oil

Moisture (wt%) 15.2 4.5–43.0 – 0.1


Solid content (wt%) 0.22 0.1–3.0 – 0.1
Dynamic viscosity (mPa s) 60 at 50°C 40–100 at 50°C 1.6–2.3 at 50°C 180 at 40°C
Density (g/mL) 1.25 0.91–1.29 0.83–0.84 0.94

pH 2.87 2.3–5.5 – –
C (wt%) 60.66 54–58 86.23–86.31 85
H (wt%) 7.70 5.5–7.0 13.14–13.27 11
N (wt%) 2.02 0–0.2 – 0.3
S (wt%) 0.15 – 0.034–0.039 –
O (wt%) 29.4 35–40 – 1.0

HHV (MJ/kg) 17.51 14–19 42.7–43.0 40

produced from the dehydration reactions during pyrolysis process. As water is miscible with
oligo-cellulose-derived compounds because of the solubilizing effect of polar hydrophilic
compounds (acids, alcohols, hydroxyaldehydes, and ketones), the presence of water in biooil
would reduce the oil viscosity and improve the flow characteristics, which are beneficial to
the combustion process. However, it may also lower the heating values of biooils, thereby
increasing the ignition delay and decreasing the combustion rate [15].
The pyrolysis biooils generally have much higher solid contents (0.1–3.0 wt%) than die-
sel oil (0 wt%) and petroleum oil (0.1 wt%). These high solid contents are mainly from the
fly chars or ashes in the condensable gases during pyrolysis process, and are of significant
importance with respect to the particulate emissions during combustion process because it
can wear the fuel system, block the filter, and clog the fuel nozzle. Generally, the larger the
particle size and the higher the particle amount, the more serious the solid content problem
is. For engine and boiler applications, the solid content of biooil should be controlled
within 1% [15].
Pyrolysis biooil may have viscosities similar to those of diesel oil and petroleum oil. Ac-
tually, the viscosity of pyrolysis biooil can vary over a wide range from as low as 10 cp to
as high as 10,000 cp depending on its temperature, and can be even higher when the biooil
is stored in poor conditions for longer periods. The viscosity of pyrolysis biooil is also affected
by water content, feedstock characteristics, process parameters, process configurations, stor-
age conditions, and storage periods. Engine companies are concerned about the viscosity of
pyrolysis biooil, because high-viscosity biooil can cause excessive fuel injection pressure dur-
ing the engine warm-up stage. On the other hand, engines would be starved for fuel at low
temperatures because the biooil would move slowly through the filters or lines due to the
high viscosities [15].

I. GENERAL
1.8 GASEOUS BIOFUELS 33
The low pH values of pyrolysis biooils are mainly attributed to the substantial amounts of
organic acids, for example, acetic acid and formic acid, which make the biooil corrosive to
common construction materials such as carbon steel and aluminum. This corrosiveness is
especially severe when the pyrolysis biooil is at elevated temperatures and increased water
contents. In that case, polyolefins are usually used as alternative construction materials due to
their resistance to the corrosiveness of biooils [15].
The pyrolysis biooils generally have higher oxygen contents (29%–40%) whereas lower car-
bon contents (54%–61%) than diesel oil and petroleum oil. As oxygen is not a combustible
element and carbon is an important combustible element, the higher oxygen contents and
lower carbon contents collectively make the pyrolysis biooils much lower in heating values.
Although the conventional electrical pyrolysis biooil and MAP biooil generally have dif-
ferent physical and chemical properties as compared to diesel oil and petroleum oil, they gen-
erally have similar physical and chemical properties. This is due to the facts that the pyrolysis
biooils are produced mainly from the depolymerization or fragmentation reactions of the
three key construction blocks of biomass (cellulose, hemicellulose, and lignin). However,
the physical and chemical properties of pyrolysis biooils may be significantly varied due
to the different pyrolysis mechanisms and parameters [109]. For instance, the lower oxygen
content (e.g., 25.78 vs 56.10 wt%) of MAP biooil led to a much higher HHV (e.g., 31.13 MJ/kg
vs 12.31 MJ/kg) than that of the conventional electrical pyrolysis biooil even when they were
obtained under the same conditions [15].
As an important renewable biofuel, pyrolysis biooil has numerous industrial usages includ-
ing: (a) combustion fuel, (b) transportation fuel, (c) power generation, (d) liquid smoke,
(e) production of diesel oil blends, (f ) production of anhydro sugars (levoglucosan),
(g) production of chemicals and resins, (h) binders for palletizing and briquetting of combustible
organic waste materials, (i) preservatives (wood preservative), (j) making adhesives, etc. [35].
Because of a number of intrinsic problems with pyrolysis biooil (e.g., high water content,
high acidity), some (further) treatments are needed to improve its quality, for example,
pretreatment of raw feedstock, introduction of reaction catalysts, optimization of operation
conditions, design of novel reactors, further treatment via hydro-processing, etc. [7, 61].

1.8 GASEOUS BIOFUELS

As compared with the raw solid biomass feedstock, gaseous biofuels generally have the
advantages of (a) being more reactive, (b) requiring less oxidants, (c) requiring simpler reac-
tors, (d) being easier to control, (e) yielding less wastes, etc. [98], and therefore gaseous
biofuels are widely studied and applied. The most important gaseous biofuels include biogas,
biohydrogen, biosyngas, etc.

1.8.1 Biogas
Biogas is also known as biomethane, which is mainly produced from anaerobic digestion
or biological fermentation of organic substances, for example, liquid manure and other
digestible biomass feedstocks [35, 37].

I. GENERAL
34 1. BIOFUELS: INTRODUCTION

Biogas can be used as a substitute for natural gas. However, there are significant differ-
ences between the characteristics of biogas and natural gas. Table 1.15 presents the compo-
sitions of natural gas and biogas. It is observed that natural gas is mainly composed of
methane (95%) and ethane (5%) whereas biogas mainly contains methane (45%–65%) and car-
bon dioxide (30%–40%). The higher content of carbon dioxide (30%–40%) contributes to the
much lower energy of the biogas compared to natural gas [1].
Still, biogas has some advantages over natural gas: (a) the raw feedstock is renewable, (b) it
is economical, (c) it does not add any GHGs to the atmosphere (the included and produced
CO2 can be fixed to form raw feedstock through photosynthesis), (d) it is produced locally
independent of foreign oil or natural gas supplies, (e) it helps in reducing the pollution pro-
duced by organic wastes, and (f ) it helps in alleviating the waste management problems [131].
Consequently, biogas is widely used as substitute for natural gas in motor vehicles, meal
cooking, house heating, and electricity generation [37, 85]. Because the raw biogas usually
contains moisture, CO2, H2S, and other impurities, purification is usually required to upgrade
the biogas before its utilization.

1.8.2 Biohydrogen
Hydrogen (pure) is a colorless and odorless gas and is the simplest and also lightest
combustible gas in nature. Table 1.16 presents the molecular weights and HHVs of some com-
monly used gases (hydrogen, carbon monoxide, methane, ethyne, ethylene, ethane, propene,
and propane). It is shown that hydrogen has the lowest molecular weight and the highest
HHV among all the combustible gases. On the other hand, when the chemical energy is re-
leased through the combustion of hydrogen, the product is nothing but water (H2O) which

TABLE 1.15 Compositions of Natural Gas and Biogas [1, 120]


Composition Natural Gas Biogas
Methane (%) 95 45–65
Ethane (%) 5 

Propane (%) + 
Butane (%) + 
Carbon dioxide (%) + 30–40
Hydrogen sulfide (%)  0.3–3
Ammonia (%)  0–1

Moisture (%)  0–10


Nitrogen (%) + 0–5
Oxygen (%)  0–2
Hydrogen (%)  0–1
+, indicate there is some content but not detailed; , indicates no.

I. GENERAL
1.8 GASEOUS BIOFUELS 35
TABLE 1.16 Properties of Some Commonly Used Gases [132–134]
Gas Formula Molecular Weight HHV (kJ/kg)
Hydrogen H2 2 141,780

Carbon monoxide CO 28 10,100


Methane CH4 16 55,510
Ethyne C2H2 26 49,910
Ethylene C2H4 28 50,300
Ethane C2H6 30 51,870
Propene C3H6 42 45,800

Propane C3H8 44 50,220

has no greenhouse effect or environmental pollution. Therefore, hydrogen is considered as an


environmentally friendly fuel and is described as the “energy carrier” of the future [135].
Biohydrogen can be broadly defined as the hydrogen produced from renewable sources.
The sources used can be energy crops, crop residues, agricultural residues, forest residues,
livestock residues, algae biomass, municipal wastes, industrial wastes, waste oil, waste water,
etc. [136, 137]. Generally, there are two main groups of methods that can be used to produce
biohydrogen: (a) thermochemical conversion, and (b) biological conversion.
Hydrogen produced from thermochemical conversion is also called biohydrogen because
the original raw feedstock is biomass. Thermochemical conversion mainly includes thermo-
chemical decomposition of water and biomass, water-gas shift reaction of (gasification or
pyrolysis) syngas, steam reforming of natural gas (or methane), steam reforming of pyrolysis
biooil, etc. [137, 138]. Although thermochemical conversion processes are cost-effective,
external energy is usually required because the reactions are mainly endothermic [139].
On the other side, hydrogen production from biological technologies generally has some
advantages: (a) the process can be performed at ambient temperature and pressure (usually in
an aqueous environment by using catalyst), (b) the process requires low or no energy, (c) the
process may be inexpensive, (d) the process is well suited for decentralized energy produc-
tion from small-scale installations in locations where biomass or wastes are available, thus
avoiding energy expenditure and costs of transportation, etc. [136, 139].
Biological conversion technologies mainly include: (1) biophotolysis, (2) photofer-
mentation, (3) dark fermentation, and (4) microbial electrolysis [104, 136, 140].
Biophotolysis uses either photosynthetic microorganisms to convert CO2 and water for
H2 production (direct biophotolysis) or photoheterophic microorganisms to convert carbo-
hydrates or organic acids to produce H2 and CO2 (indirect biophotolysis) [138]. Photo-
fermentation uses photofermentative bacteria to photodecompose organic compounds
[136, 140]. Dark fermentation produces hydrogen using anaerobic or facultative anaerobic
bacteria [136, 141], and microbial electrolysis converts electron equivalents in organic com-
pounds to H2 by combining microbial metabolism with bioelectrochemical reactions in
microbial electrolysis cell [142]. Currently, each biological conversion technology has its tech-
nical disadvantages or challenges, for example, biophotolysis and photofermentation are

I. GENERAL
36 1. BIOFUELS: INTRODUCTION

light dependent; accumulation of acidogenic by-products may inhibit dark fermentation; mi-
crobial electrolysis may cause cell lysis when the undissociated soluble metabolites permeate
the cell membrane of bacteria, etc. Generally, dark fermentation (a) is relatively less energy
intensive, (b) is technically simpler, (c) requires moderate conditions, (d) requires lower op-
eration costs, and (e) is more stable and robust. Therefore it appears to be more feasible and
practical than the other biological conversion technologies [135, 140, 142, 143].

1.8.3 Biosyngas
Different from biogas and biohydrogen, biosyngas is mainly the target product of biomass
gasification (sometimes, pyrolysis can also produce high-quality syngas). Biosyngas is a mix-
ture of mainly CO, H2, N2, CO2, and hydrocarbons (e.g., CH4, C2H4, and C2H6). Very small
amounts of moisture, NH3, H2S, and tars may also exist.
Based on the gasification agent used, biomass gasification can be air gasification (using air
[98]), oxygen gasification (using oxygen [99]), steam gasification (using steam [100]), carbon
dioxide gasification (using carbon dioxide [101]), supercritical water gasification (using
supercritical water [102]), etc. Because air is cheap and readily available, and also the gasifi-
cation process and reactor are relatively simple, air gasification is the most widely studied
and used among these approaches mentioned above. Because the gasification agent (air) con-
tains a high content of N2 (79 vol%), the produced syngas would also have a very high N2
content (e.g., >60 vol% [144]), which drastically dilutes the biosyngas.
To improve the quality (heating value) of biosyngas, oxygen, steam, carbon dioxide, or su-
percritical water can be used instead of air. Among these gasification agents, oxygen and
steam are more commonly used.
The essence of using pure oxygen is mainly to reduce the dilution effect of N2, thereby
increasing the volume fractions of the other gases. As presented in Table 1.17 oxygen gasifi-
cation nearly increases the volume fractions of all the gases as compared with air gasification.
Because most of the gas components are high-quality (high heating value) gases, the heating
value of biosyngas has significantly increased (7.7–9.5 MJ/Nm3 vs 4.4–5.2 MJ/Nm3). How-
ever, the introduced pure oxygen may also react with (or burn) the combustible gas compo-
nents, thereby decreasing the quality of biosyngas. Also, the other barrier is the high cost of
pure oxygen production.
The essence of steam gasification is that (a) it avoids the dilution effect of N2, and (b) it
significantly increases the contents of H2 and CO (mainly through water-gas shift reaction).
Therefore, the quality of biosyngas has significantly improved (as depicted in Table 1.17).
However, external heat is needed to generate and heat the steam. Also, the tar content of
biosyngas can be very high.
The compositions and properties of syngas obtained from electrical pyrolysis and micro-
wave pysolysis are also illustrated in Table 1.17. As compared with biomass gasification
including air gasification, oxygen gasification, and steam gasification, both electrical pyroly-
sis and microwave pysolysis generally result in higher quality biosyngas HHVs. However,
the pyrolysis approaches generally yield lower gas yields and higher tar yields.
As compared with the raw solid biofuel (or feedstock), the resultant biosyngas (or syngas)
is more versatile, for example, it can be directly used as natural gas, gas turbine fuel, boiler

I. GENERAL
1.9 CONCLUSIONS AND PERSPECTIVES 37
TABLE 1.17 Compositions and Properties of Syngas
Item Gasification Pyrolysis
a b b b
Air Steam Oxygen Electrical Microwavec
Feedstock Rice husk Sawdust Sawdust Sawdust Corn stover
Reaction temperature (°C) 850–900 800–1200 800–1200 800–1200 900
H2 (vol%) 2.5–3.9 18.7–48.0 10.0–26.2 15.7–47.1 22.5–23.6
CO (vol%) 19.1–21.8 26.7–55.5 29.5–47.8 46.2–58.2 30.8–38.8

CO2 (vol%) 12.4–13.2 5.5–19.8 24.4–49.2 3.2–6.0 16.6–20.6


CH4 (vol%) 2.6–3.2 3.8–12.9 1.6–5.6 2.2–13.9 7.4–10.6
C2H2 (vol%) 0.08–0.12 0–2.5 0–1.8 0.04–2.5
C2H4 (vol%) 1.0–1.3 0–6.0 0.1–2.6 0.05–6.2
C2H6 (vol%) 0.07–0.17
N2 (vol%) 57.2–61.4
3
Tar content (g/Nm ) 3.2–4.0
Tar yield (wt%) 0–7.7 0–0.7 0–11.4 7.0–14.7
3 d d d
LHV (MJ/Nm ) 4.4–5.2 10.4–17.2 7.7–9.5 11.7–17.7 8.9–11.2
a
Ref. [144].
b
Ref. [145].
c
Ref. [146].
d
Calculated.

fuel, engine fuel, and fuel cell feed [85, 97]. The produced biosyngas can be upgraded to en-
rich the hydrogen content through water-gas-shift reaction [35]. It can be converted to provide
a wide range of long carbon chain biofuels, such as synthetic diesel, aviation fuel, or methanol
via the gas-to-liquid platform [61]. It can also be processed to produce liquid fuels or hydro-
carbons known as Fisher-Tropsch oil (FT oil) or green motor fuel through FT synthesis pro-
cesses [35]. Also, it can be used to produce lubricants as well as chemical commodities [97].
On the other hand, the included tar is also a high energy content biofuel, which can be
combusted in burners and boilers to produce heat, mechanical power, and electricity [97].

1.9 CONCLUSIONS AND PERSPECTIVES

Biofuels are abundantly available, environmentally friendly, renewable, biodegradable,


and sustainable. The definitions of biofuel implicate that a biofuel feedstock is also a biofuel,
and can be converted to other types of biofuels through proper technologies. Based on the
feedstock sources and conversion technologies, biofuels can be classified into five groups:
zeroth-generation biofuels, first-generation biofuels, second-generation biofuels, third-
generation biofuels, and fourth-generation biofuels. The zeroth-generation biofuels are
mainly the natural biofuels or raw feedstocks which mainly include oil crops, lignocellulosic

I. GENERAL
38 1. BIOFUELS: INTRODUCTION

biomass, solid wastes, and algae. If these feedstocks are treated by certain technologies such
as physical technologies (e.g., physical extraction, distillation, etc.), chemical technologies
(e.g., solvent extraction, supercritical extraction, supercritical water conversion, etc.), thermo-
chemical technologies (e.g., combustion, gasification, pyrolysis, liquidation, etc.), and bio-
chemical technologies (e.g., digestion, fermentation, hydrolysis, interesterification,
transesterification, etc.), heat, work, electricity, and/or resultant biofuels can be obtained.
The resultant biofuels can be in the form of solid (e.g., solid biomass, biochar), liquid
(e.g., bioethanol, biodiesel, and biooil), or gas (e.g., biogas, biohydrogen, and biosyngas).
The diversities of biofuel feedstocks and production technologies render the biofuels with
a wide range of properties and applications, and each technology comes with both advan-
tages and disadvantages. Advanced drop-in biofuel production technologies is still of signif-
icant importance in the future studies. Environmentally friendly and cost-effective biofuels
are the constant pursuit.

Acknowledgments
Acknowledgments are attributed to China Scholarship Council (No. 201506125122), National Natural Science Foun-
dation of China (No. 51606048), National Natural Science Foundation of China (No. 21766019), The Key Research and
Development Program of Jiangxi Province (No. 20171BBF60023), Minnesota’s Environment and Natural Resources
Trust Fund (ENRTF) through the processes of the Legislative-Citizen Commission on Minnesota Resources
(LCCMR), and University of Minnesota Center for Biorefining.

References
[1] M. Guo, W. Song, J. Buhain, Bioenergy and biofuels: history, status, and perspective, Renew. Sust. Energ. Rev.
42 (2015) 712–725.
[2] M.F. Demirbas, M. Balat, H. Balat, Potential contribution of biomass to the sustainable energy development,
Energy Convers. Manag. 50 (2009) 1746–1760.
[3] E. Sadeghinezhad, S.N. Kazi, F. Sadeghinejad, A. Badarudin, M. Mehrali, R. Sadri, M.R. Safaei,
A comprehensive literature review of bio-fuel performance in internal combustion engine and relevant costs
involvement, Renew. Sust. Energ. Rev. 30 (2014) 29–44.
[4] C.H.B. Cruz, G.M. Souza, L.A.B. Cortez, Biofuels for transport, in: Chapter 11 in Future Energy, second ed.,
Elsevier Ltd. London, UK, 2014, pp. 215–244
[5] A.L. Demain, M.A. Báez-Vásquez, Chapter 14: Biofuels of the present and the future, in: New and Future De-
velopments in Catalysis, Elsevier, Amsterdam, The Netherlands, 2013, pp. 325–370.
[6] M. Tabatabaei, A. Mollahosseini, G. Najafpour, Biofuel production, in: Chapter 20 in Biochemical Engineering
and Biotechnology, second ed., Elsevier, Amsterdam, The Netherlands, 2015, pp. 597–630.
[7] C. Du, X. Zhao, D. Liu, C.S.K. Lin, K. Wilson, R. Luque, J. Clark, Introduction: an overview of biofuels and pro-
duction technologies, in: Chaper 1 in Handbook of Biofuels Production, second ed., Elsevier, Duxford, UK,
2016, pp. 3–12.
[8] I.L. Garcı́a, Feedstocks and challenges to biofuel development, in: Chapter 5 in Handbook of Biofuels Produc-
tion, second ed., Elsevier, Duxford, UK, 2016, pp. 85–118.
[9] Y. Zhang, A.E. Ghaly, B. Li, Determination of the exergy of four wheat straws, Am. J. Biochem. Biotechnol. 9 (3)
(2013) 338–347.
[10] Y. Zhang, A.E. Ghaly, B. Li, Comprehensive investigation into the exergy values of six rice husks, Am. J. Eng.
Appl. Sci. 6 (2) (2013) 216–225.
[11] Y. Zhang, A.E. Ghaly, B. Li, Availability and physical properties of residues from major agricultural crops for
energy conversion through thermochemical processes, Am. J. Agric. Biol. Sci. 7 (3) (2012) 312–321.
[12] Y. Zhang, X. Yu, B. Li, H. Zhang, J. Liang, Y. Li, Exergy characteristics of woody biomass, Energy Sources, Part A
38 (16) (2016) 2438–2446.

I. GENERAL
REFERENCES 39
[13] Y. Zhang, Q. Wang, B. Li, H. Li, W. Zhao, Is there a general relationship between the exergy and HHV for rice
residues? Renew Energy 117 (2018) 37–45.
[14] T. Suganya, M. Varman, H.H. Masjuki, S. Renganathan, Macroalgae and microalgae as a potential source for
commercial applications along with biofuels production: a biorefinery approach, Renew. Sust. Energ. Rev.
55 (2016) 909–941.
[15] Y. Zhang, P. Chen, S. Liu, L. Fan, N. Zhou, M. Min, Y. Cheng, P. Peng, E. Anderson, Y. Wang, Y. Wan, Y. Liu,
B. Li, R. Ruan, Microwave-assisted pyrolysis of biomass for bio-oil production, in: Chapter 6 in Pyrolysis,
InTech, Rijeka, Croatia, 2017, pp. 129–166.
[16] T.L. Chew, S. Bhatia, Catalytic processes towards the production of biofuels in a palm oil and oil palm biomass-
based biorefinery, Bioresour. Technol. 99 (2008) 7911–7922.
[17] Y. Zhang, W. Zhao, B. Li, H. Li, Understanding the sustainability of fuel from the viewpoint of exergy, Eur. J.
Sustain. Dev. Res. 2 (1) (2018) 09.
[18] A. Alaswad, M. Dassisti, T. Prescott, A.G. Olabi, Technologies and developments of third generation biofuel
production, Renew. Sust. Energ. Rev. 51 (2015) 1446–1460.
[19] D. Russo, M. Dassisti, V. Lawlor, A.G. Olabi, State of the art of biofuels from pure plant oil, Renew. Sust. Energ.
Rev. 16 (2012) 4056–4070.
[20] A. Pradhan, C. Mbohwa, Development of biofuels in South Africa: challenges and opportunities, Renew. Sust.
Energ. Rev. 39 (2014) 1089–1100.
[21] B. Ghobadian, Liquid biofuels potential and outlook in Iran, Renew. Sust. Energ. Rev. 16 (2012) 4379–4384.
[22] Y. Su, P. Zhang, Y. Su, An overview of biofuels policies and industrialization in the major biofuel producing
countries, Renew. Sust. Energ. Rev. 50 (2015) 991–1003.
[23] Webster, Biofuel, Definition of Biofuel, https://www.merriam-webster.com/dictionary/biofuel, 2018.
[24] B.A. Robertson, P.J. Doran, Biofuels and biodiversity: the implications of energy sprawl, in: Encyclopedia of
Biodiversity, second ed., Academic Press, Waltham, MA, 2013, pp. 528–539.
[25] A.B. Delshad, L. Raymond, V. Sawicki, D.T. Wegener, Public attitudes toward political and technological
options for biofuels, Energy Policy 38 (2010) 3414–3425.
[26] M.F. Demirbas, M. Balat, H. Balat, Biowastes-to-biofuels, Energy Convers. Manag. 52 (2011) 1815–1828.
[27] Food and Agriculture Organization of the United Nations (FAO), Introducing the International Bioenergy Plat-
form (IBEP), FAO, Rome, 2006.
[28] F. Saladini, N. Patrizi, F.M. Pulselli, N. Marchettini, S. Bastianoni, Guidelines for emergy evaluation of first,
second and third generation biofuels, Renew. Sust. Energ. Rev. 66 (2016) 221–227.
[29] IEA, Biofuel Roadmap, http://www.ieabioenergy.com/wp-content/uploads/2013/10/IEA-Biofuel-Roadmap.
pdf, 2011.
[30] S.C. Bhatia, Biofuels: a review, in: Chapter 16 in Advanced Renewable Energy Systems, Woodhead Publishing
India Pvt. Ltd., Daryaganj, New Delhi, India, 2014, pp. 403–425
[31] A. Demirbas, Progress and recent trends in biofuels, Prog. Energy Combust. Sci. 33 (2007) 1–18.
[32] A.A. Kiss, A.C. Dimian, G. Rothenberg, Solid acid catalysts for biodiesel production—towards sustainable
energy, Adv. Synth. Catal. 348 (2006) 75–81.
[33] L.A. Schulte, T.A. Ontl, L. GLD, Biofuels and biodiversity, wildlife habitat restoration, in: Encyclopedia of
Biodiversity, second ed., 2013, pp. 540–551.
[34] J. Hill, Life cycle analysis of biofuels, in: Encyclopedia of Biodiversity, second ed., 2013, pp. 627–630.
[35] S.N. Naik, V.V. Goud, P.K. Rout, A.K. Dalai, Production of first and second generation biofuels: a comprehen-
sive review, Renew. Sust. Energ. Rev. 14 (2) (2010) 578–597.
[36] S. Akhlaghi, U.W. Gedde, M.S. Hedenqvist, M.T.C. Braña, M. Bellander, Deterioration of automotive rubbers in
liquid biofuels: a review, Renew. Sust. Energ. Rev. 43 (2015) 1238–1248.
[37] P. Morone, L. Cottoni, Biofuels: technology, economics, and policy issues, in: Chaper 4 in Handbook of Biofuels
Production, second ed., Elsevier, Duxford, UK, 2016, pp. 61–83.
[38] IEA, Energy Technology Perspective—Scenario and Strategies to 2050, IEA, Paris, 2008.
[39] V.K. Gupta, R. Potumarthi, A. O’Donovan, C.P. Kubicek, G.D. Sharma, M.G. Tuohy, Bioenergy research: an
overview on technological developments and bioresources, in: Chapter 2 in Bioenergy Research: Advances
and Applications, Elsevier, Waltham, MA, 2014, pp. 23–47.
[40] L.D. Zhu, E. Hiltunen, E. Antila, J.J. Zhong, Z.H. Yuan, Z.M. Wang, Microalgal biofuels: flexible bioenergies for
sustainable development, Renew. Sust. Energ. Rev. 30 (2014) 1035–1046.
[41] C.J. Franco, S. Zapata, I. Dyner, Simulation for assessing the liberalization of biofuels, Renew. Sust. Energ. Rev.
41 (2015) 298–307.

I. GENERAL
40 1. BIOFUELS: INTRODUCTION

[42] J. Milano, H.C. Ong, H.H. Masjuki, W.T. Chong, M.K. Lam, P.K. Loh, V. Vellayan, Microalgae biofuels as an
alternative to fossil fuel for power generation, Renew. Sust. Energ. Rev. 58 (2016) 180–197.
[43] G. Joshi, J.K. Pandey, S. Rana, D.S. Rawat, Challenges and opportunities for the application of biofuel, Renew.
Sust. Energ. Rev. 79 (2017) 850–866.
[44] M.A. Scaife, A. Merkx-Jacques, D.L. Woodhall, R.E. Armenta, Algal biofuels in Canada: status and potential,
Renew. Sust. Energ. Rev. 44 (2015) 620–642.
[45] M.K. Lam, K.T. Lee, Scale-up and commercialization of algal cultivation and biofuel production, in: Chapter 12
in Biofuels From Algae, Elsevier, Burlington, MA, 2014, pp. 261–286.
[46] M.K. Lam, K.T. Lee, Bioethanol production from microalgae, in: Chapter 12 in Handbook of Marine Microalgae,
Academic Press, London, UK, 2015, pp. 197–208.
[47] S.J.M. Algayyim, A.P. Wandel, T. Yusaf, I. Hamawand, Production and application of ABE as a biofuel, Renew.
Sust. Energ. Rev. 82 (2018) 1195–1214.
[48] M.Y. Noraini, H.C. Ong, M.J. Badrul, W.T. Chong, A review on potential enzymatic reaction for biofuel pro-
duction from algae, Renew. Sust. Energ. Rev. 39 (2014) 24–34.
[49] H. Hao, Z. Liu, F. Zhao, J. Ren, S. Chang, K. Rong, J. Du, Biofuel for vehicle use in China: current status, future
potential and policy implications, Renew. Sust. Energ. Rev. 82 (2018) 645–653.
[50] K.T. Lee, S. Lim, Y.L. Pang, H.C. Ong, W.T. Chong, Integration of reactive extraction with supercritical fluids for
process intensification of biodiesel production: prospects and recent advances, Prog. Energy Combust. Sci.
45 (2014) 54–78.
[51] T. Bhaskar, A. Pandey, Advances in thermochemical conversion of biomass–introduction, in: Chapter 1 in Recent
Advances in Thermo-Chemical Conversion of Biomass, Elsevier, Amsterdam, The Netherlands, 2015, pp. 3–30.
[52] P.A. Fokaides, E. Christoforou, Life cycle sustainability assessment of biofuels, in: Chaper 3 in Handbook of
Biofuels Production, second ed., Woodhead Publishing, Duxford, UK, 2016, pp. 41–60.
[53] A.M.N. Renzaho, J.K. Kamara, M. Toole, Biofuel production and its impact on food security in low and middle
income countries: implications for the post-2015 sustainable development goals, Renew. Sust. Energ. Rev.
78 (2017) 503–516.
[54] M. Acheampong, F.C. Ertem, B. Kappler, P. Neubauer, In pursuit of sustainable development goal (SDG) num-
ber 7: will biofuels be reliable? Renew. Sust. Energ. Rev. 75 (2017) 927–937.
[55] A. Taghizadeh-Alisaraei, S.H. Hosseini, B. Ghobadian, A. Motevali, Biofuel production from citrus wastes: a
feasibility study in Iran, Renew. Sust. Energ. Rev. 69 (2017) 1100–1112.
[56] X. Li, X. Luo, Y. Jin, J. Li, H. Zhang, A. Zhang, J. Xie, Heterogeneous sulfur-free hydrodeoxygenation catalysts
for selectively upgrading the renewable bio-oils to second generation biofuels, Renew. Sust. Energ. Rev.
82 (2018) 3762–3797.
[57] W.H. Liew, M.H. Hassim, D.K.S. Ng, Review of evolution, technology and sustainability assessments of biofuel
production, J. Clean. Prod. 71 (2014) 11–29.
[58] E.K. Vakkilainen, Solid biofuels and combustion, in: Chaper 2 in Steam Generation From Biomass, Butterworth-
Heinemann, Oxford, UK, 2017, pp. 18–56.
[59] M.H. Duku, S. Gu, E.B. Hagan, A comprehensive review of biomass resources and biofuels potential in Ghana,
Renew. Sust. Energ. Rev. 15 (2011) 404–415.
[60] C.L. Williams, A. Dahiya, P. Porter, Introduction to bioenergy, in: Chapter 1 in Bioenergy, Academic Press,
London, UK, 2015, pp. 5–36.
[61] D.P. Ho, H.H. Ngo, W. Guo, A mini review on renewable sources for biofuel, Bioresour. Technol. 169 (2014)
742–749.
[62] A.K. Azad, M.G. Rasul, M.M.K. Khan, S.C. Sharma, M.A. Hazrat, Prospect of biofuels as an alternative transport
fuel in Australia, Renew. Sust. Energ. Rev. 43 (2015) 331–351.
[63] K.R. Jegannathan, E.S. Chan, P. Ravindra, Harnessing biofuels: a global renaissance in energy production? Re-
new. Sust. Energ. Rev. 13 (2009) 2163–2168.
[64] S.P. Singh, D. Singh, Biodiesel production through the use of different sources and characterization of oils and
their esters as the substitute of diesel: a review, Renew. Sust. Energ. Rev. 14 (1) (2010) 200–216.
[65] J. Du, P. Liu, Z.H. Liu, D.G. Sun, C.Y. Tao, Fast pyrolysis of biomass for bio-oil with ionic liquid and microwave
irradiation, J. Fuel Chem. Technol. 38 (5) (2010) 554–559.
[66] Y.F. Huang, P.T. Chiueh, W.H. Kuan, S.L. Lo, Microwave pyrolysis of rice straw: products, mechanism, and
kinetics, Bioresour. Technol. 142 (2013) 620–624.
[67] Y.F. Huang, P.T. Chiueh, S.L. Lo, A review on microwave pyrolysis of lignocellulosic biomass, Sustainable En-
viron. Res. 26 (3) (2016) 103–109.

I. GENERAL
REFERENCES 41
[68] X. Wang, W. Morrison, Z. Du, Y. Wan, X. Lin, P. Chen, R. Ruan, Biomass temperature profile development and
its implications under the microwave assisted pyrolysis condition, Appl. Energy 99 (2012) 386–392.
[69] Y.F. Huang, W.H. Kuan, C.C. Chang, Y.M. Tzou, Catalytic and atmospheric effects on microwave pyrolysis of
corn stover, Bioresour. Technol. 131 (2013) 274–280.
[70] B.J. Lin, W.H. Chen, Sugarcane bagasse pyrolysis in a carbon dioxide atmosphere with conventional and
microwave-assisted heating, Front. Energy Res. 3 (4) (2015) 1–9.
[71] Y. Wan, P. Chen, B. Zhang, C. Yang, Y. Liu, X. Lin, R. Ruan, Microwave-assisted pyrolysis of biomass: catalysts
to improve product selectivity, J. Anal. Appl. Pyrolysis 86 (1) (2009) 161–167.
[72] A.A. Salema, F.N. Ani, Pyrolysis of oil palm empty fruit bunch biomass pellets using multimode microwave
irradiation, Bioresour. Technol. 125 (2012) 102–107.
[73] R. Omar, A. Idris, R. Yunus, K. Khalid, M.I.A. Isma, Characterization of empty fruit bunch for microwave-
assisted pyrolysis, Fuel 90 (4) (2011) 1536–1544.
[74] O. Mašek, V. Budarin, M. Gronnow, K. Crombie, P. Brownsort, E. Fitzpatrick, P. Hurst, Microwave and slow
pyrolysis biochar-comparison of physical and functional properties, J. Anal. Appl. Pyrolysis 100 (2013) 41–48.
[75] T. V€ais€
anen, A. Haapala, R. Lappalainen, L. Tomppo, Utilization of agricultural and forest industry waste and
residues in natural fiber-polymer composites: a review, Waste Manag. 54 (2016) 62–73.
[76] H.Y. Li, B. Wang, J.L. Wen, X.F. Cao, S.N. Sun, R.C. Sun, Availability of four energy crops assessing by the en-
zymatic hydrolysis and structural features of lignin before and after hydrothermal treatment, Energy Convers.
Manag. 155 (2018) 58–67.
[77] K.C. Surendra, R. Ogoshi, H.M. Zaleski, A.G. Hashimoto, S.K. Khanal, High yielding tropical energy crops for
bioenergy production: effects of plant components, harvest years and locations on biomass composition,
Bioresour. Technol. 251 (2018) 218–229.
[78] M.K. Lam, K.T. Lee, Microalgae biofuels: a critical review of issues, problems and the way forward, Biotechnol.
Adv. 30 (2012) 673–690.
[79] A. Raheem, W.A.K.G.W. Azlina, Y.H.T. Yap, M.K. Danquah, R. Harun, Thermochemical conversion of
microalgal biomass for biofuel production, Renew. Sust. Energ. Rev. 49 (2015) 990–999.
[80] H. Chen, D. Zhou, G. Luo, S. Zhang, J. Chen, Macroalgae for biofuels production: progress and perspectives,
Renew. Sust. Energ. Rev. 47 (2015) 427–437.
[81] A. Demirbas, M.F. Demirbas, Importance of algae oil as a source of biodiesel, Energy Convers. Manag. 52 (1)
(2011) 163–170.
[82] M.E. Edeseyi, A.Y. Kaita, R. Harun, M.K. Danquah, C. Acquah, J.K.M. Sia, Rethinking sustainable biofuel mar-
keting to titivate commercial interests, Renew. Sust. Energ. Rev. 52 (2015) 781–792.
[83] R.P. Rastogi, A. Pandey, C. Larroche, D. Madamwar, Algal green energy—R&D and technological perspectives
for biodiesel production, Renew. Sust. Energ. Rev. 82 (2018) 2946–2969.
[84] M. Kumar, S. Sundaram, E. Gnansounou, C. Larroche, I.S. Thakur, Carbon dioxide capture, storage and pro-
duction of biofuel and biomaterials by bacteria: a review, Bioresour. Technol. 247 (2018) 1059–1068.
[85] P. Champagne, Biomass, in: Chapter 9 in Future Energy, Elsevier Ltd., London, UK, 2008, pp. 151–170
[86] R. Dutta, U. Sarkar, A. Mukherjee, Extraction of oil from Crotalaria Juncea seeds in a modified Soxhlet appa-
ratus: physical and chemical characterization of a prospective bio-fuel, Fuel 116 (2014) 794–802.
[87] S. Rahman, R. Helleur, S. MacQuarrie, S. Papari, K. Hawboldt, Upgrading and isolation of low molecular
weight compounds from bark and softwood bio-oils through vacuum distillation, Sep. Purif. Technol.
194 (2018) 123–129.
[88] G.N. Mezza, A.V. Borgarello, N.R. Grosso, H. Fernandez, M.C. Pramparo, M.F. Gayol, Antioxidant activity of
rosemary essential oil fractions obtained by molecular distillation and their effect on oxidative stability of sun-
flower oil, Food Chem. 242 (2018) 9–15.
[89] Y. Zhang, A.E. Ghaly, B. Li, Physical properties of wheat straw varieties cultivated under different climatic and
soil conditions in three continents, Am. J. Eng. Appl. Sci. 5 (2) (2012) 98–106.
[90] L. Jiang, J. Liang, X. Yuan, H. Li, C. Li, Z. Xiao, H. Huang, H. Wang, G. Zeng, Co-pelletization of sewage sludge
and biomass: the density and hardness of pellet, Bioresour. Technol. 166 (2014) 435–443.
[91] K.T. Tan, K.T. Lee, Biodiesel production in supercritical fluids, in: Chapter 14 in Biofuels, Academic Press,
Oxford, UK, 2011, pp. 339–352.
[92] O.V. Okoro, Z. Sun, J. Birch, Meat processing waste as a potential feedstock for biochemicals and biofuels—a
review of possible conversion technologies, J. Clean. Prod. 142 (2017) 1583–1608.
[93] H.B. Goyal, D. Seal, R.C. Saxena, Bio-fuels from thermochemical conversion of renewable resources: a review,
Renew. Sust. Energ. Rev. 12 (2008) 504–517.

I. GENERAL
42 1. BIOFUELS: INTRODUCTION

[94] E. Suali, R. Sarbatly, Conversion of microalgae to biofuel, Renew. Sust. Energ. Rev. 16 (2012) 4316–4342.
[95] Y. Zhang, X. Fan, B. Li, H. Li, X. Gao, Assessing the potential environmental impact of fuel using exergy-cases of
wheat straw and coal, Int. J. Exergy 23 (1) (2017) 85–100.
[96] Y. Zhang, X. Gao, B. Li, H. Li, W. Zhao, Assessing the potential environmental impact of woody biomass using
quantitative universal exergy, J. Clean. Prod. 176 (2018) 693–703.
[97] Y. Zhang, Y. Zhao, X. Gao, B. Li, J. Huang, Energy and exergy analyses of syngas produced from rice husk
gasification in an entrained flow reactor, J. Clean. Prod. 95 (2015) 273–280.
[98] Y. Zhang, Y. Zhao, B. Li, X. Gao, B. Jiang, Energy and exergy characteristics of syngas produced from air gas-
ification of walnut sawdust in an entrained flow reactor, Int. J. Exergy 23 (3) (2017) 244–262.
[99] M. Niu, Y. Huang, B. Jin, X. Wang, Oxygen gasification of municipal solid waste in a fixed-bed gasifier, Chin. J.
Chem. Eng. 22 (2014) 1021–1026.
[100] Y. Zhang, B. Li, H. Li, B. Zhang, Exergy analysis of biomass utilization via steam gasification and partial ox-
idation, Thermochim. Acta 538 (2012) 21–28.
[101] X. Gao, Y. Zhang, B. Li, Y. Zhao, B. Jiang, Determination of the intrinsic reactivities for carbon dioxide gasifi-
cation of rice husk chars through using random pore model, Bioresour. Technol. 218 (2016) 1073–1081.
[102] L. Guo, H. Jin, Y. Lu, Supercritical water gasification research and development in China, J. Supercrit. Fluids
96 (2015) 144–150.
[103] X. Gao, Y. Zhang, B. Li, X. Yu, Model development for biomass gasification in an entrained flow gasifier using
intrinsic reaction rate submodel, Energy Convers. Manag. 108 (2016) 120–131.
[104] IEA, Main reactions during biomass gasification, http://www.ieatask33.org/app/webroot/files/file/
various/Main%20reactions%20during%20biomass%20gasification.pdf, 2015.
[105] Y. Zhang, B. Li, H. Li, H. Liu, Thermodynamic evaluation of biomass gasification with air in autothermal gas-
ifiers, Thermochim. Acta 519 (1–2) (2011) 65–71.
[106] M. Balat, M. Balat, K. Elif, H. Balat, Main routes for the thermo-conversion of biomass into fuels and chemicals.
Part 2: gasification systems, Energy Convers. Manag. 50 (12) (2009) 3158–3168.
[107] T. Damartzis, A. Zabaniotou, Thermochemical conversion of biomass to second generation biofuels through
integrated process design—a review, Renew. Sust. Energ. Rev. 15 (2011) 366–378.
[108] P. Manara, A. Zabaniotou, Towards sewage sludge based biofuels via thermochemical conversion—a review,
Renew. Sust. Energ. Rev. 16 (2012) 2566–2582.
[109] Y. Zhang, P. Chen, S. Liu, P. Peng, M. Min, Y. Cheng, E. Anderson, N. Zhou, L. Fan, C. Liu, G. Chen, Y. Liu,
H. Lei, B. Li, R. Ruan, Effects of feedstock characteristics on microwave-assisted pyrolysis—a review, Bioresour.
Technol. 230 (2017) 143–151.
[110] P. Jayasinghe, K. Hawboldt, A review of bio-oils from waste biomass: focus on fish processing waste, Renew.
Sust. Energ. Rev. 16 (2012) 798–821.
[111] K. Azizi, M.K. Moraveji, H.A. Najafabadi, A review on bio-fuel production from microalgal biomass by using
pyrolysis method, Renew. Sust. Energ. Rev. 82 (2018) 3046–3059.
[112] A. Taghizadeh-Alisaraei, H.A. Assar, B. Ghobadian, A. Motevali, Potential of biofuel production from pistachio
waste in Iran, Renew. Sust. Energ. Rev. 72 (2017) 510–522.
[113] R.W. Jenkins, A.D. Sutton, D.J. Robichaud, Pyrolysis of biomass for aviation fuel, in: Chapter 8 in Biofuels for
Aviation, Academic Press, London, UK, 2016, pp. 191–215.
[114] S. Liu, Y. Zhang, L. Fan, N. Zhou, G. Tian, X. Zhu, Y. Cheng, Y. Wang, Y. Liu, P. Chen, R. Ruan, Bio-oil pro-
duction from sequential two-step catalytic fast microwave-assisted biomass pyrolysis, Fuel 196 (2017) 261–268.
[115] L. Fan, Y. Zhang, S. Liu, N. Zhou, P. Chen, Y. Cheng, M. Addy, Q. Lu, M.M. Omar, Y. Liu, Y. Wang, L. Dai,
E. Anderson, P. Peng, H. Lei, R. Ruan, Bio-oil from fast pyrolysis of lignin: effects of process and upgrading
parameters, Bioresour. Technol. 241 (2017) 1118–1126.
[116] Y. Zhang, S. Liu, L. Fan, N. Zhou, M.M. Omar, P. Peng, E. Anderson, M. Addy, Y. Cheng, Y. Liu, B. Li, J. Snyder,
P. Chen, R. Ruan, Oil production from microwave-assisted pyrolysis of a low rank American brown coal, En-
ergy Convers. Manag. 159 (2018) 76–84.
[117] Z.M.A. Bundhoo, Microwave-assisted conversion of biomass and waste materials to biofuels, Renew. Sust.
Energ. Rev. 82 (2018) 1149–1177.
[118] S.A. Razzak, M.M. Hossain, R.A. Lucky, A.S. Bassi, H. de Lasa, Integrated CO2 capture, wastewater treatment
and biofuel production by microalgae culturing—a review, Renew. Sust. Energ. Rev. 27 (2013) 622–653.
[119] V. Parisutham, T.H. Kim, S.K. Lee, Feasibilities of consolidated bioprocessing microbes: from pretreatment to
biofuel production, Bioresour. Technol. 161 (2014) 431–440.

I. GENERAL
REFERENCES 43
[120] M.C. Gould, Bioenergy and anaerobic digestion, in: Chapter 18 in Bioenergy, Academic Press, London, UK,
2015, pp. 297–317.
[121] P. Binod, K.U. Janu, R. Sindhu, A. Pandey, Hydrolysis of lignocellulosic biomass for bioethanol production,
in: Chapter 10 in Biofuels, Academic Press, Oxford, UK, 2011, pp. 229–250.
[122] P.F. Stanbury, A. Whitaker, S.J. Hall, An introduction to fermentation processes, in: Chapter 1 in Principles of
Fermentation Technology, third ed., 2017, pp. 1–20.
[123] Fermentation, From Wikipedia, the Free Encyclopedia, https://en.wikipedia.org/wiki/Fermentation, 2018.
[124] C.R. Soccol, V. Faraco, S. Karp, L.P.S. Vandenberghe, V. Thomaz-Soccol, A. Woiciechowski, A. Pandey, Ligno-
cellulosic bioethanol: current status and future perspectives, in: Chapter 5 in Biofuels, 2011, pp. 101–122.
[125] H.S. Kambo, A. Dutta, A comparative review of biochar and hydrochar in terms of production, physico-
chemical properties and applications, Renew. Sust. Energ. Rev. 45 (2015) 359–378.
[126] J.C. Escobar, E.S. Lora, O.J. Venturini, E.E. Yáñez, E.F. Castillo, O. Almazan, Biofuels: environment, technology
and food security, Renew. Sust. Energ. Rev. 13 (2009) 1275–1287.
[127] F. Y€uksel, B. Y€uksel, The use of ethanol–gasoline blend as a fuel in an SI engine, Renew. Energy 29 (7) (2004)
1181–1191.
[128] A. Sanjid, H.H. Masjuki, M.A. Kalam, S.M.A. Rahman, M.J. Abedin, S.M. Palash, Impact of palm, mustard,
waste cooking oil and Calophyllum inophyllum biofuels on performance and emission of CI engine, Renew.
Sust. Energ. Rev. 27 (2013) 664–682.
[129] G.T. Ang, K.T. Tan, K.T. Lee, Recent development and economic analysis of glycerol-free processes via super-
critical fluid transesterification for biodiesel production, Renew. Sust. Energ. Rev. 31 (2014) 61–70.
[130] B.F. Towler, Ethanol, biodiesel, and biomass, in: Chapter 12 in the future of energy, 2014, pp. 257–271.
[131] R. Chandra, H. Takeuchi, T. Hasegawa, Methane production from lignocellulosic agricultural crop wastes: a
review in context to second generation of biofuel production, Renew. Sust. Energ. Rev. 16 (2012) 1462–1476.
[132] Y. Zhang, Thermodynamic Study on Gasification Process of Biomass Fuels (in Chinese), Dissertation for the
Doctoral Degree in Engineering, Harbin Institute of Technology, Harbin, China, 2012.
[133] HyARC, Lower and Higher Heating Values of Fuels, https://www.h2tools.org/hyarc/calculator-tools/lower-
and-higher-heating-values-fuels, 2018.
[134] R.N. Walters, S.M. Hackett, R.E. Lyon, Heats of Combustion of High Temperature Polymers, http://large.
stanford.edu/publications/coal/references/docs/hoc.pdf, 2018.
[135] A. Noblecourt, G. Christophe, C. Larroche, P. Fontanille, Hydrogen production by dark fermentation from pre-
fermented depackaging food wastes, Bioresour. Technol. 247 (2018) 864–870.
[136] G.D. Saratale, R.G. Saratale, J.S. Chang, Biohydrogen from renewable resources, in: Chapter 9 in Biohydrogen,
2013, pp. 185–221.
[137] W. Khetkorn, R.P. Rastogi, A. Incharoensakdi, P. Lindblad, D. Madamwar, A. Pandey, C. Larroche, Microalgal
hydrogen production—a review, Bioresour. Technol. 243 (2017) 1194–1206.
[138] T. Bhaskar, B. Balagurumurthy, R. Singh, M.K. Poddar, Thermochemical route for biohydrogen production,
in: Chapter 12 in biohydrogen, 2013, pp. 285–316.
[139] M.K. Lam, K.T. Lee, Biohydrogen production from algae, in: Chapter 8 in biohydrogen, 2013, pp. 161–184.
[140] P.C. Hallenbeck, Fundamentals of biohydrogen, in: Chapter 2 in Biohydrogen, 2013, pp. 25–43.
[141] Z. Trad, J. Akimbomi, C. Vial, C. Larroche, M.J. Taherzadeh, J.P. Fontaine, Development of a submerged an-
aerobic membrane bioreactor for concurrent extraction of volatile fatty acids and biohydrogen production,
Bioresour. Technol. 196 (2015) 290–300.
[142] S.V. Mohan, A. Pandey, Biohydrogen production: an introduction, in: Chapter 1 in Biohydrogen, 2013, pp. 1–24.
[143] Z. Trad, J.P. Fontaine, C. Larroche, C. Vial, Multiscale mixing analysis and modeling of biohydrogen production
by dark fermentation, Renew. Energy 98 (2016) 264–282.
[144] Y. Zhao, D. Feng, Z. Zhang, S. Sun, X. Zhou, J. Luan, J. Wu, Experimental study of cyclone pyrolysis—
suspended combustion air gasification of biomass, Bioresour. Technol. 243 (2017) 1241–1246.
[145] Y. Zhang, S. Kajitani, M. Ashizawa, Y. Oki, Tar destruction and coke formation during rapid pyrolysis and gas-
ification of biomass in a drop-tube furnace, Fuel 89 (2010) 302–309.
[146] Q. Xie, F.C. Borges, Y. Cheng, Y. Wan, Y. Li, X. Lin, Y. Liu, F. Hussain, P. Chen, R. Ruan, Fast microwave-
assisted catalytic gasification of biomass for syngas production and tar removal, Bioresour. Technol.
156 (2014) 291–296.

I. GENERAL

You might also like