Open access peer-reviewed chapter

Woody Feedstock Pretreatments to Enhance Pyrolysis Bio-oil Quality and Produce Transportation Fuel

Written By

Hamid Rezaei, Fahimeh Yazdanpanah, Jim Choon Lim, Anthony Lau and Shahab Sokhansanj

Submitted: 22 September 2018 Reviewed: 02 October 2018 Published: 24 December 2018

DOI: 10.5772/intechopen.81818

From the Edited Volume

Biomass for Bioenergy - Recent Trends and Future Challenges

Edited by Abd El-Fatah Abomohra

Chapter metrics overview

1,382 Chapter Downloads

View Full Metrics

Abstract

Lignocellulosic biomass as a potential renewable source of energy has a near-zero CO2 emission. Pyrolysis converts biomass to a liquid fuel and increases the energy density and transportability. The pyrolysis bio-oil shows promising properties to substitute the conventional fossil fuels. But, unprocessed biomass is low in bulk and energy density; high in moisture; heterogeneous in physical and chemical properties, highly hygroscopic and difficult to handle. That is why the biomass needs mechanical, chemical and/or thermal pretreatments to turn into a more homogeneous feedstock and minimize the post-treatment fuel upgrading. This chapter explains the effects that various pretreatments such as size reduction, drying, washing and thermal pretreatments have on the quality and quantity of bio-oil. Washing with water or acid/alkali solutions extracts the minerals that consequently reduces the ash and shortens the reactor clean-out cycle. Torrefaction is gaining attention as an effective pretreatment to modify the quality of biomass in terms of physical and chemical properties. Torrefaction produces a uniform biomass with lower moisture, acidity and oxygen contents and higher energy density and grindability than raw biomass. Pyrolysis of torrefied biomass produces bio-oil with enhanced compositional and physical properties such as a higher heating value and increased C (lower O/C ratio).

Keywords

  • pyrolysis
  • pretreatment
  • torrefaction
  • washing
  • bio-oil
  • bio-fuel

1. Introduction

Crude oil and coal with high energy density and stable properties to store and transport are preferred fuels for industrial and home applications [1]. The recent reduction in affordable fossil fuels and environmental concerns increased greenhouse gas (GHG) emissions from these fuels that have motivated the public and private sectors to develop renewable energy sources. In recent years, the governments of Canada and the USA adopted regulations to phase out coal-fired power plants by 2030 and investigate other resources to reduce GHG emission. Biomass feedstocks are considered as the renewable resource with a near-zero CO2 input-output balance [2, 3, 4]. Biomass is the only renewable energy source which may be used in all solid, liquid and gas phases [5]. But compared to fossil fuels, biomass is lower in energy (17–19 MJ/kg) and bulk density (60–100 kg/m3) [6]; heterogeneous in physical, chemical and thermal properties; high in moisture [6], mineral [7] and oxygen contents [8]; highly hygroscopic [9] and difficult to handle [6]. Converting biomass to secondary liquid or gaseous fuels through thermal conversion processes is a way of increasing the energy density and transportability. Mechanical, chemical and thermal pretreatments are able to modify the biomass properties in order to produce a more homogeneous fuel and minimize the post-treatment fuel upgrading.

The current article reviews the feedstock properties that are important for the fast pyrolysis process and the potential pretreatments to modify the biomass properties are explained. Specifically, drying, grinding, washing and torrefaction processes and their influence on the final bio-oil product are thoroughly explained.

Advertisement

2. Pyrolysis

Pyrolysis is a thermal process to convert biomass particles to secondary solid, liquid and/or gaseous fuels [10]. The liquid bio-oil obtained from pyrolysis of biomass shows promising properties to substitute the conventional fossil fuels. The breakdown of biomass particles takes place in a pyrolysis reactor by exposing them to an oxygen-free or deficit environment. As the temperature of the particle increases, different stages of thermal treatment occur. Drying happens first when a moist particle is heated up to temperatures of about 150°C [11]. At temperatures of 200–300°C and in the absence of oxygen/air, various degrees of torrefaction happen and release volatile [12, 13, 14]. At temperatures higher than 300°C, more severe decomposition of biomass happens and structural changes such as a change in chemical composition and porosity of particles occur in the solid matrix.

Pyrolysis processes are generally categorized as “slow” and “fast” according to the time taken for processing the feed into pyrolysis products. Slow pyrolysis utilizes low temperatures of 300–400°C over a residence time of 30 minutes to hours to maximize char formation. Fast pyrolysis is a rapid thermal decomposition process to maximize the liquid fraction of products. In industrial fast pyrolysis, biomass particles would be exposed suddenly to an environment of 400–600°C to heat up rapidly. The combined effects of moisture content, particle size [15, 16, 17] and reactor temperature [18, 19] make a non-linear temperature profile inside the particle which affects the rate of conversion. The residence time of the volatile fraction in the pyrolysis reactor and its contact with solid particles should be less than 2–3 seconds to prevent the secondary reactions. Pyrolysis takes place in a complete absence of oxygen. The volatiles are quickly removed and quenched to maximize liquid yield. Figure 1 displays a schematic diagram of an industrial pyrolysis process. In the following sections, a comprehensive literature review is presented to discuss the effects of various parameters related to feedstock and pre-treatments on the efficiency of the pyrolysis process.

Figure 1.

Schematic diagram of the pyrolysis process, including drying, grinding, pyrolyzer, separation stages and condenser.

2.1 Feedstock composition

A plant-based dry biomass has the main fractions of cellulose, hemicelluloses, lignin, extractives and minerals. Wood extractives, or wood extracts, are low molecular weight molecules that are extracted from wood using solvents or other extraction methods. The extractives are the waxes, fatty acids, resin acids and terpenes of a tree. Table 1 lists the composition of a broad range of biomass species. The composition of biomass changes the chemical composition and heating value of pyrolysis products [33]. Higher lignin content reduces the bio-oil yield at the expense of higher biochar yield in the pyrolysis process [29]. The extractives have the highest heating value in the woody biomass. The heating values of biomass components are 17–18, 16–17, 25–26 and 33–38 GJ/tonne for cellulose, hemicellulose, lignin and extractives, respectively [34].

Biomass species Composition (wt.%) Refs. Ultimate analysis (wt.%) Ash (wt.%) HHV (MJ/kg) Refs.
Cel.1 Hem.2 Lig.3 C H O
Pine wood 42.1 27.7 25.0 [20] 45.9 5.3 48.2 0.35 18.98 [21]
Beech wood 41.7 37.1 18.9 [22] 46.7 5.7 47.6 1.04 [22]
Douglas fir 50.0 17.8 28.3 [23] 50.7 6.0 42.5 0.50 20.07 [24]
Spruce wood 41.1 20.9 28.0 [20] 48.9 6.0 44.6 0.30 19.26 [20]
Poplar aspen 49.9 22.4 18.1 [25]
Poplar wood 42.2 16.6 25.6 [26] 48.4 5.8 43.7 1.43 19.71 [26]
Birch wood 35.7 25.1 19.3 [20] 49.0 6.3 44.1 0.30 18.40 [27]
Corn stover 36.4 22.6 16.6 [28]
Corn cob 52.0 32.5 15.5 [29] 49.0 5.6 43.8 1.10 [29]
Wheat straw 38.2 24.7 23.4 [28] 40.7 5.8 52.9 10.58 16.24 [30]
Rice straw 34.2 24.5 23.4 [31] 36.9 5.0 37.9 11.70 16.78 [31]
Rice straw 32.1 26.5 12.5 [32]
Olive husk 25.2 24.2 50.6 [29] 50.2 6.4 38.4 4.10 [29]
Tea waste 33.2 23.3 43.5 [29] 48.2 5.5 44.3 1.50 [29]

Table 1.

Composition and ultimate analysis, ash content and calorific value of various biomass feedstock.

Cellulose.


Hemicellulose.


Lignin.


Each sub-component of the biomass has a specific thermal reactivity. For example, cellulose decomposes at 300–400°C; hemicellulose decomposes at 200–300°C and lignin decompose continuously in the wide temperature range of 180–600°C [35]. This specification directly influences the temperature range at which the material thermally decomposes and consequently the optimum pyrolysis temperature at which the maximum bio-oil yield is obtained. Table 2 lists pyrolysis liquid yield of various biomass species at tested temperatures and depicts the variability of produced bio-oil yield among the pyrolysis of different feedstocks. Apart from variability in conversion rate among species, the conversion rate increases with the reaction temperature.

Biomass type Pyrolysis temperature (°C) Bio-oil yield (wt.%) Refs.
Pine wood 450 55.0 [36]
Waste furniture sawdust 450 65.0 [37]
Wood sawdust 650 74.0 [38]
Corncob 550 56.8 [39]
Municipal, livestock and wood waste 500 39.7 [40]
Pine wood 450 50 [41]
Rice husks 450 60 [42]
Corn cobs and corn stover 650 61.6 [43]
Sugar cane waste 470 56.5 [44]

Table 2.

Bio-oil yield in fast pyrolysis of different biomass feedstocks.

In addition to the chemical composition, the elemental and proximate analysis of the biomass feedstock changes the properties of the produced fuel. From the elemental point of view, the biomass is mainly contained of carbon, hydrogen and oxygen (Table 1). The composition of elements affects the storage properties of the liquid fuel. Compared to conventional fossil fuels, biomass has a high amount of ~40–50% oxygen content. High oxygen content makes the produced liquid fuel unstable, corrosive and consequently not qualified for transportation and storage [16].

The minor elements present in the biomass material are minerals such as potassium, chlorine, sulfur, silicon, calcium and magnesium [45]. Minerals are present in all biomass species, in a much lower amount than carbon, hydrogen and oxygen elements. Agricultural biomass has much more mineral contents than woody biomass. In a thermal process, minerals turn to ash (Table 1). The pyrolysis biochar typically contains up to 90% of the biomass minerals [46]. Ash shifts the size distribution of the char to smaller sizes that make their recovery from the gas stream challenging. An incomplete separation of char and volatiles causes continuous secondary reactions in the liquid phase [16, 47, 48] that accelerates the aging phenomenon and contribute to its instability [49, 50]. Aging phenomena is defined as a slow increase in viscosity bio-oil resulting from secondary reactions [16]. Minerals have a catalytic effect on the rate of secondary reactions [51].

2.2 Pyrolysis products

Products of the pyrolytic reaction are biochar and hot volatile that need to be separated and cooled down quickly. The residence time of char and volatile in direct contact should be less than 2–3 seconds [5, 15, 52]. Char is greatly porous and has a catalytic effect on volatile cracking to non-condensable gases, more char and also non-desirable volatiles [1]. The main product of fast pyrolysis is liquid bio-oil that is obtained by condensation of volatiles, yielding up to 75 wt.% based on a dry feed basis [15]. Table 3 lists a few key properties of bio-oil and the properties of petroleum grade fuel (Fuel No. 6). Bio-oil has a heating value of about half of conventional fossil fuel oil, 17–19 MJ/kg for bio-oil versus 35–40 MJ/kg for petroleum fossil fuel [52].

Characteristics Fast pyrolysis bio-oil Heavy petroleum fuel
Wet Dry
Water content (wt.%) 15–25 0.1
Insoluble solids (%) 0.5–0.8 0.01
Carbon (%) 39.5 55.8 85.2
Hydrogen (%) 7.5 6.1 11.1
Oxygen (%) 52.6 37.9 1.0
Nitrogen (%) <0.1 0.3
Sulfur (%) <0.05 2.3
Ash (%) 0.2–0.3 <0.1
Heating value (MJ/kg) 17 40
Density (g/ml) 1.23 0.94
Viscosity (cp) @ 50°C 10–150 180

Table 3.

Properties of wood-derived bio-oil and heavy petroleum fuel [53].

The final application of produced bio-oil depends on its quality and properties such as solid content, stability during storage and transportation, oxygen and water contents and its pH value [54]. High oxygen and water content make the bio-oil less stable and less competitive with conventional fossil fuels. Low stabilization, including phase separation and polymerization, and acidic and corrosive properties of bio-oil make its storage and transportation difficult [39, 55]. Based on National Renewable Energy Laboratory’s (NREL) recommendation, Multi-Year Program Plan (MYPP 2011), the bio-oil must maintain its consistency for at least 6 months of storage. Based on the published studies, a drastic reduction in water content of bio-oil, insoluble solids and oxygen content and a substantial increase in the calorific value of bio-oil are needed for upgrading to fuel grade.

Gas and biochar are generated as by-products of the pyrolysis process. Optimizing their properties helps to find an application for by-products to make the overall process more economically feasible. They have the potential to provide the heat required for the pyrolysis process. The non-condensable gases of the pyrolysis process have a substantial amount of CO2 and CH4. Combustion of these permanent gases makes heat for the pyrolysis process. Biochar has various interesting properties. First one is that the biochar also has a high calorific value. The energy density for charcoal is 9–11 GJ/m3, while energy density for wood and coal are 8–10 and 25–40 GJ/m3, respectively [1]. Second, biochar is enriched with carbon that is a good soil amendment. In addition, adding char to soil moves the carbon from atmosphere to soil [56]. Third but not last, pyrolysis biochar has promising adsorption properties for heavy metals [57], dyes [58, 59] and aromatics [60] in agricultural and industrial effluents. The biomass properties influence the biochar properties to be used for adsorption purposes. For example, woody biomass contains more lignin content than agricultural biomass and its biochar has higher surface area than other biomass-derived biochars. Then, it is speculated to have better adsorptivity properties [61].

Advertisement

3. Feedstock pretreatment

Physical properties like particle size, particle shape and density influence the material handling and flowability of particles that are leading properties for an un-interrupted feeding system [62]. Chemical properties like elemental and proximate analyses influence the pyrolysis products’ distribution and properties. Although biomass has inherent heterogeneous properties [1], various physical, chemical and thermal pretreatments contribute to achieving homogeneity in properties of biomass. Homogeneous feedstock with uniform physical and chemical properties plays a key role in a pyrolysis process and quality of its products. Although pretreatments create an additional feedstock cost, they facilitate feeding a wide range of biomass species with broad properties and are beneficial due to producing uniform and homogeneous feedstock for power plants.

3.1 Moisture reduction

Biomass particles are dried for a more-efficient thermal combustion in pyrolysis reactors to produce bio-oil [63, 64]. A fresh biomass has a high moisture content of up to 80% [34]. High moisture content reduces the heating value of the fuel and shifts the ignition point to higher temperatures [6], and inhibits the rise of temperature inside the particles and conversion reactor [65, 66]. Reduced particle and reactor temperatures diminish the liquid yield at the expense of a larger fraction of biochar and non-condensable gases [22, 66]. Water has a catalytic effect on volatile cracking. Di Blasi [67] showed that moisture content conducts the pyrolysis reaction to a low activation energy path that promotes the formation of char, non-condensable gases and more water. Biomass should be dried down to less than 10% moisture content to improve the quality of produced fuel [15]. Pyrolysis of dry material produces less water compared to wet biomass. Bio-oil with a lower amount of water has higher heating value, a lower ignition point, higher combustion rate and a lower potential of phase separation [6].

Rezaei [68] stated that size of particles [69], initial moisture content [70], drying temperature [69, 71], relative humidity of drying gas [72, 73] and particle heating rate [74] influence the rate of moisture loss. Rezaei et al. [75] showed that biomass shrinks during the moisture loss that influences the drying rate. Dehydration of fresh wood causes a reduction in the dimension of wood in a direction normal to the microfibril orientation, whereas the longitudinal shrinkage is usually negligible [76]. Mazzanti et al. [77] showed that the longitudinal shrinkage for poplar wood is negligible. Taylor et al. [78] showed that the radial shrinkage of beech wood is about 70 times of longitudinal shrinkage.

3.2 Size reduction

Grinding reduces the dimensions of a particle and increases the particle’s specific surface area (ratio of the particle’s surface area to its mass). The same relationship holds for a bulk of particles where the surface area of the solids increases in a given volume of bulk particles. Therefore, the smaller particles have a more exposed surface to the raised-temperature environment to boost the rate of heat and mass transfer between the particles and surrounding. Rezaei et al. [79, 80] showed that larger particles (from 1 to 5 mm) have a delay in heat and mass transfer and reduced rate of pyrolysis (Figure 2). On the other hand, grinding the biomass to smaller particles requires more energy input. Rezaei et al. [81] reported that grinding wood chips to produce particles with average sizes of 1, 1.8, 3 and 5 mm requires about 124.0, 85.7, 52.6 and 28.2 kJ/kg, respectively. Van Der Stelt et al. [82] measured the power required for size reduction of coal, torrefied woodcutting, willow and demolition wood. He showed that to have particles smaller than 0.6 mm, the grinding power consumption increases sharply. The balance of the conversion rate and energy consumption identifies the optimum particle size appropriate for a fast pyrolysis process.

Figure 2.

Effect of particle size on the pyrolysis rate of ground wood particles at 500°C [80].

Figure 3 shows the published data on yields of three phases of solid, liquid and gas in pyrolysis of a range of particle size. In the range of particle size up to 2 mm, only the apricot stone pyrolyzed at 800°C showed a yield that was sensitive to particles size. An increase in particle size increased the char yield at the expense of less liquid yield [15, 22, 26, 83]. Some researchers did not observe any effect of particle size on the yield of products. For example, Şensöz et al. [84, 85] did not observe any meaningful influence of particle size on the pyrolysis products of rapeseed (Brassica napus L.) in the range of 0.22–0.85 mm [84] and of debarked pine in the range of 2–5 mm [85]. Encinar et al. [86] reported that liquid yield was independent of biomass particle size during pyrolysis of 0.4–2.0 mm grape residue and olive residue particles pyrolyzed at 500°C. On the other hand, some researchers showed that larger particles in pyrolysis decrease liquid yield and increase biochar yield [22, 83, 87]. NikAzar et al. [22] found out that size increase from 53 to 63 μm to 270–500 μm declined liquid yield from 53 to 38%. He claimed the particle’s core temperature diminished by particle size and caused the change in the yield of products. This result was confirmed by some other researchers [15, 16, 17, 26, 29, 88].

Figure 3.

Effect of biomass particle size on the yield of products in a fast pyrolysis process.

Most of the published data recommend that particles in the range of 1–2 mm are appropriate for fast pyrolysis [17, 52, 82, 89]. However, the effect of particle size on the pyrolysis process needs more work. The range of particle sizes tested in various studies is seldom comparable and a study on a wider range of particle size seems necessary. In addition, the challenges associated with commercial grinding and feeding the bulky lightweight biomass in the reactor should be taken into account that is out of the scope of the current chapter.

3.3 De-mineralization

The mineral elements in biomass may be listed mostly as potassium, chlorine, sulfur, silicon, calcium and magnesium [45]. The mineral content of biomass exists on the particle surface because of contact with soil during harvest and/or transportation, or within the material as biogenic characteristics.

The minerals of the biomass present as ash or fouling in conversion vessels. In a pyrolysis process, the biochar typically contains up to 90% of the biomass minerals [46] that changes the physical properties of char. Ash shifts the size distribution of the char to smaller sizes that hardens its full separation from produced volatile. A partial char separation results in a high solid content bio-oil that cannot be used as turbine fuel [49]. Furthermore, minerals act as a catalyst and cause continuous secondary reactions in liquid phase [16, 47, 48] at the expense of char formation [5190]. More secondary reactions promote an increases in bio-oil viscosity with time and accelerate aging phenomenon [50]. Mineral contamination accelerates the catalytic breaking of levoglucosan into unwanted hydroxyl acetaldehyde compounds [51, 91]. Former is desired part of volatiles and latter is an undesirable portion of volatiles. Addition of 0.05 wt.% NaCl to an ash-free cellulose decreases the levoglucosan formation yield by a factor of 6 [92].

One efficient pretreatment to reduce the mineral content is washing the biomass with water, acidic and/or alkaline solutions. Washing with dilute acid and hot water results in a slight decomposition of hemicellulose [93]. Washing with dilute alkali disrupts the lignin structure and solubilize the hemicellulose [94]. Washing biomass prior to pyrolysis takes away a huge amount of minerals from the biomass, up to 70% of the initial minerals [1, 17].

Das et al. [95] studied the effect of various washing solutions, concentrations and the time of washing on ash content and products’ yields of sugarcane bagasse pyrolysis (Figure 4). The ash content decreased from 1.83% before washing to 0.03%. The only washing solution that had a reverse effect was 5 M HCl. It must have been due to the adding chlorine element into the biomass. The important point is the effect of washing on yield of total liquid versus the bio-oil. The total liquid contains bio-oil and aqueous solutions. All washing solutions reduced the total liquid yield but boosted the bio-oil yield.

Figure 4.

Effect of de-mineralization pretreatment on biomass ash content and products’ yield of sugarcane bagasse pyrolysis [95].

Table 4 lists different washing solutions based on demineralization yield, ash content, char and liquid yield and maximum decomposition rate [96]. Washing demineralized the biomass up to 98%, enhanced liquid yield, lowered the char yield, shifted up the decomposition rate and reduced the low molecular weight compounds in pyrolysis products. Solution temperature also influences the demineralization yield. Deng et al. [7] showed that demineralization of candlenut wood raised from 8% at 30°C to 35% to 90°C. Mineral removal increased the higher heating value from 16.53 to 17.82 MJ/kg and the rate of devolatilization increased too. Natural rain and season of raining change the composition of minerals in biomass [97].

Washing solution Demineralization yield (%) Char yield (%) Volatiles yield (wt.%) Max. decomposition rate (wt.%/°C) Tp3 (°C)
No washing 14.9 85.1 0.96 362
HCl1 69.3 11.8 88.2 1.55 366
HF2 97.3 10.2 89.8 1.15 368
Deionized water 97.7 11.0 89.0 1.23 372
Tap water 98.2 11.4 88.6 1.19 376

Table 4.

Specifications for pyrolysis of demineralized poplar wood at 550˚C [96].

Hydrochloric acid.


Hydrofluoric acid.


Temperature at which maximum decomposition rate happens.


3.4 Torrefaction

Torrefaction is a mild thermal treatment that modifies the structure and chemical composition of biomass by removing hemicelluloses [98], dehydrating and partially reducing cellulose and lignin [99]. Li et al. [13] described torrefaction as a mild heat treatment of biomass at a temperature range of 200–300°C, prolonging 15–30 minutes. Westover et al. [6] divided torrefaction into three stages of non-reactive drying (50–150°C), reactive drying (150–200°C) and destructive stage (200–300°C). During thermal treatment, biomass releases moisture up to the temperature of 150–170°C [11]. In 180–270°C, exothermic hemicellulose degradation happens and biomass turns to brown color. At this stage, torrefaction reaction produces more water, CO2, acetic acid and phenols [12, 13, 14, 35]. The released gases are combustible and may be used to provide a portion of process energy. Beyond 270°C, the reactions are more exothermic and produce CO and some other heavier products such as CH4 and C2H6. Torrefaction continues to a temperature of 300°C, where pyrolysis starts.

Torrefied biomass contains about 80% of the mass and 90% of the energy of the initial biomass [82, 99]. Similar to other thermal processes, mass and energy yield of torrefaction depends on biomass species [99], particle size [14, 26], operating temperature [6, 13, 33, 100, 101] and residence time [13, 33, 101]. Biomass torrefaction has been recognized as a feasible technique to convert raw biomass to a high energy density, hydrophobic, grindable, homogeneous, low moisture content (<5%) [6], better storage stability [34] and low-oxygen-content fuel that is a suitable feedstock for pyrolysis [102].

Torrefaction changes the stiffness and glass transition temperature (Tg) of biomass [6] that contributes to the grindability of the material. Ground torrefied biomass particles are smaller, drier and more uniform in size and physical and chemical properties [13, 14]. Table 5 lists the specific grinding energy of raw and torrefied pine and logging residue. Higher torrefaction temperature makes the biomass structure more brittle and reduces the specific grinding energy. Specific grinding energy consumption reduced by 90% from 240 kWh/t for raw pine chips to 24 kWh/t for torrefied pine chips at 300°C.

Sample Specific grinding energy (kWh/t) Hemicel. (wt.%) Sample Specific grinding energy (kWh/t) Hemicel. (wt.%)
Untreated-PC1 237.7 15.19 Untreated-LR3 236.7 14.77
TPC2-225°C 102.6 12.87 TLR4-225°C 113.8 13.26
TPC-250°C 71.4 6.94 TLR-250°C 110.4 5.87
TPC-275°C 52.0 0.99 TLR-275°C 78.0 5.23
TPC-300°C 23.9 0.56 TLR-300°C 37.6 1.04

Table 5.

Specific energy consumption for grinding of untreated and torrefied biomass with a residence time of 30 minutes [98].

Pine chips.


Torrefied pine chips.


Logging residues.


Torrefied logging residues.


Using the torrefied biomass as a feedstock for fast pyrolysis has various benefits. The level of these modifications depends on torrefaction temperature and residence time. First of all, torrefaction removes the hemicellulose and enriches the biomass into the cellulose and lignin (Figure 5). Presence of hemicellulose intensively decreases the yield of levoglucosan and promotes the formation of hydroxyl acetaldehyde. The torrefied feedstock may eliminate or reduce this interaction as it removes hemicelluloses.

Figure 5.

Composition of raw and torrefied pine at 240 and 280°C [103].

Bio-oil of torrefied biomass has a lower water content that increases bio-oil’s stability (less phase separation) [100, 104, 105], lower oxygen/carbon ratio and higher calorific value [104] compared to the bio-oil produced from non-torrefied biomass [5, 89, 99, 101]. Because torrefaction extracts contain acidic condensable volatiles such as acetic acid, furfural, formic acid, methanol, lactic acid and phenol from the biomass [106], the bio-oil has a lower acidity and aldehydes content [17, 105]. Klinger [105] pyrolyzed the torrefied material and observed 27% water reduction, 36% CO reduction, 55% CO2 reduction and 67% acetic acid reduction in the produced volatiles compared to volatiles obtained from non-torrefied biomass.

Despite all benefits, pyrolysis of torrefied biomass has a lower yield of total liquid. Liaw et al. [100] reported that pyrolysis total liquid yields of raw Douglas fir and torrefied Douglas fir at 280, 320 and 370°C was 59, 55, 48 and 30%, respectively [100]. Boateng [99] and Zheng [5, 104] reported that more sever torrefaction decreases the total liquid yield and increases the yield of biochar and permanent gases. Tables 6 and 7 list the results of Zheng’s work for pyrolysis of raw and torrefied corncob.

Properties and compounds Raw corncob Torrefied (240°C) Torrefied (300°C)
Water content (wt.%) 35.00 30.00 21.00
High heating value (MJ/kg) 14.85 16.49 17.21
pH 2.68 3.30 3.34
Kinematic viscosity @ 20°C (cSt) 3.42 7.27 12.62
Acids (95% acetic acid) (%wb) 7.16 5.65 4.75
Ketones (wt.%) 8.16 7.42 6.02
Furans (wt.%) 0.98 0.8 0.76
Phenols (wt.%) 2.45 4.51 5.27

Table 6.

Physical properties and a few main chemical contents of bio-oil produced from pyrolysis (at 500°C) of raw and torrefied corncob at 240 and 300°C for 20 minutes [104].

Properties Raw corncob Residence time: 20 minutes Temperature: 275°C
250°C 275°C 300°C 10 minutes 60 minutes
Ultimate analysis of torrefied corncob (wt.%)
C 43.87 45.10 48.54 57.86 47.64 57.33
H 6.06 6.20 6.65 6.75 6.10 5.93
O 49.50 48.01 44.06 34.43 45.61 35.92
N 0.53 0.61 0.70 0.86 0.60 0.75
S 0.04 0.08 0.05 0.11 0.04 0.06
O/C 1.13 1.06 0.91 0.60 0.96 0.63
Yield of pyrolysis products (wt.%)
Bio-oil 57.20 55.15 47.60 40.74 50.36 39.35
Biochar 21.14 24.57 30.21 38.19 26.31 40.70
Non-condensable gas 21.66 20.28 22.19 21.07 23.33 19.95
Physical properties of bio-oil
Water content 35.0 33.0 25.0 21.0 30.0 22.0
Higher heating value (MJ/kg) 14.85 15.12 16.49 17.21 15.58 17.09
pH 2.68 2.97 3.30 3.34 2.88 3.35
Kinematic viscosity @ 20°C (cSt) 3.42 3.96 7.27 12.62 4.23 12.48
Crystallinity (%) 20.06 25.74 34.82 27.35 23.26 22.83

Table 7.

Effect of temperature and residence time of corncob torrefaction on ultimate analysis of feedstock, the yield of pyrolysis products (at 500°C) and physical properties of bio-oil [5].

Water content reduced from 35% in bio-oil produced from raw corncob to 21% in bio-oil produced from torrefied corncob at 300°C. The pH increased in the bio-oil prepared from torrefied corncob. The viscosity of bio-oil also increased probably due to a reduction in water content. The acetic acid content decreased moderately with increasing torrefaction temperature and residence time. The furfural content also decreased gradually with torrefaction temperature and residence time. Acetic acid and furfural are mainly derived from hemicellulose and cellulose. Ren et al. [101, 107] conducted a research on Douglas fir in two sequent articles. They determined the effects of pyrolysis temperature and torrefaction time, as a pretreatment, on the characterization of produced total bio-oil and syngas. Total bio-oil is the sum of condensed liquids from torrefaction and pyrolysis. Figure 6 shows that bio-oil yield from raw (untreated) biomass ranged from 31 [107] to 53% [101] for pyrolysis at 400 and 450°C, respectively. The bio-oil yield from a 450°C pyrolysis decreased to 51 and 46% for 8 and 15 minutes torrefaction pretreatment at 275°C was carried out. Longer torrefaction reduced slightly the yield of bio-oil. Torrefaction also altered the compositions of syngas by reducing CO2 and increasing H2 and CH4. The syngas produced from pyrolysis step was rich in H2, CH4 and CO implying that the syngas quality was significantly improved by torrefaction process [105]. The quantity of syngas increased with the severity of the torrefaction.

Figure 6.

Bio-oil yield from raw and torrefied Douglas fir at the various duration and temperature [101, 107].

Table 8 summarizes a qualitative analysis of benefits of feedstock thermal pretreatment on logistical and quality of bio-oil properties. Column 1 lists major properties that a feedstock would gain as a result of thermal pretreatment. The degree of change in properties depends upon the severity of thermal treatment. Column 2 lists the benefits of these properties on improving conversion efficiency or the quality of produced bio-oil when the pretreated feedstock is pyrolyzed. Column 3 outlines the benefits of feedstock properties from a logistical perspective, that is, handling, transport, storage, unit operations like blending and feeding the feedstock to the pyrolysis reactor. Finally, column 4 lists the potential monetary benefits of thermal pretreatments mostly due to improvements in logistics.

Feedstock properties Conversion and quality of bio-oil benefit Logistics benefit Economic benefit
Grindability Uniform small size particles for high heat transfer in the reactor Torrefied particles are dry and a lower tendency for electrostatic charged Lower investment and operating cost in grinder and grinder operation
Hydrophobicity Less water content due to a decrease in OH and COOH groups Store unprotected, exposure to rain, long shelf life Lower storage cost, lower transport cost
Low moisture content Reduced water in the bio-oil Reduced mass to handle, stable in storage Expensive drying is not required
High heat value Bio-oil will have a higher heat value due to increased C and low O/C ratio Can be blended with coal and other high heat value products Lower cost ($/GJ)
No need to design a new combustion chamber
Homogeneity Predictable conversion performance in the pyrolysis reactor Reduced quality control, may easily become a commodity Lower management cost
High density (after grinding) Can be controlled to a precise particle size and density Improved flowability, low off-gas emissions Reduced cost of shipping and storage
Thermal degradation Lower acidity Higher energy density (GJ/m3) Higher $/GJ
Wet fractionation (wet torrefaction) Less chlorine and ash Access to low-quality feedstock (e.g. bark) Reduced overall feedstock cost ($/t)

Table 8.

Overview of benefits of torrefied feedstock for bio-oil production.

These improvements are in comparison with the untreated feedstock.

Advertisement

4. Conclusions

The energy produced from lignocellulosic biomass is renewable, environmentally friendly and carbon-neutral. Characteristics of bio-oil produced in fast pyrolysis showed that it is capable to substitute conventional petroleum-based fuels. This chapter explained the fast pyrolysis process and the pretreatments that contribute to acquiring a transportation bio-oil with enhanced physical and chemical properties. The modification of biomass feedstock properties through available pretreatment techniques plays a key role in the pyrolysis process. Pretreatments change the biomass feedstock’s size, shape, mineral content, composition, hygroscopic properties, homogeneity, grindability, stability and transportability.

Torrefaction as a mild pyrolysis of biomass feedstock is a promising pretreatment that modifies the structure and chemical composition of biomass by removing hemicelluloses, dehydrating and enriching biomass in cellulose and lignin. Torrefaction makes the structure of biomass more brittle that make it more grindable. Grinding torrefied biomass produces more uniform particles with smaller and narrower particle size distribution. Pyrolysis of torrefied material enhances the properties of produced bio-oil. This enhancement is consistent with the DOE MYPP’s target that “additional processing of the bio-oil is required to enable it to become a feedstock suitable for use in a petroleum refinery at several entry points.” However, this is penalized with a liquid yield reduction.

With all the opportunities of renewable biomass energy, some important issues are still needed to be solved to commercialize the process. To find the least-cost combination of torrefaction along with biomass fast pyrolysis, further testing of the two along with economic analysis are required.

References

  1. 1. Kasparbauer RD. The effects of biomass pretreatments on the products of fast pyrolysis [master of science dissertation]. Mechanical Engineering, Iowa State University; 2009
  2. 2. Carlson TR, Tompsett GA, Conner WC, Huber GW. Aromatic production from catalytic fast pyrolysis of biomass-derived feedstocks. Topics in Catalysis. 2009;52:241-252. DOI: 10.1007/s11244-008-9160-6
  3. 3. Biagini E, Narducci P, Tognotti L. Size and structural characterization of lignin-cellulosic fuels after the rapid devolatilization. Fuel. 2008;87:177-186. DOI: 10.1016/j.fuel.2007.04.010
  4. 4. Lu H, Ip E, Scott J, Foster P, Vickers M, Baxter LL. Effects of particle shape and size on devolatilization of biomass particle. Fuel. 2010;89:1156-1168. DOI: 10.1016/j.fuel.2008.10.023
  5. 5. Zheng A, Zhao Z, Chang S, Huang Z, Wang X, He F, et al. Effect of torrefaction on structure and fast pyrolysis behavior of corncobs. Bioresource Technology. 2013;128:370-377. DOI: 10.1016/j.biortech.2012.10.067
  6. 6. Westover TL, Phanphanich M, Clark ML, Rowe SR, Egan SE, Zacher AH, et al. Impact of thermal pretreatment on the fast pyrolysis conversion of southern pine. Biofuels. 2013;4:45-61
  7. 7. Deng L, Zhang T, Che D. Effect of water washing on fuel properties, pyrolysis and combustion characteristics, and ash fusibility of biomass. Fuel Processing Technology. 2013;106:712-720. DOI: 10.1016/j.fuproc.2012.10.006
  8. 8. Wang Y, He T, Liu K, Wu J, Fang Y. From biomass to advanced bio-fuel by catalytic pyrolysis/hydro-processing: Hydrodeoxygenation of bio-oil derived from biomass catalytic pyrolysis. Bioresource Technology. 2012;108:280-284. DOI: 10.1016/j.biortech.2011.12.132
  9. 9. Zanzi R, Sjöström K, Björnbom E. Rapid pyrolysis of straw at high temperature. In: Bridgwater AV, Boocock DGB, editors. Developments in Thermochemical Biomass Conversion. The Netherlands: Springer; 1997. pp. 61-66. DOI: 10.1007/978-94-009-1559-6_4; 10.1007/978-94-009-1559-6_4
  10. 10. Bridgwater AV, Meier D, Radlein D. An overview of fast pyrolysis of biomass. Organic Geochemistry. 1999;30:1479-1493. DOI: 10.1016/S0146-6380(99)00120-5
  11. 11. Vlaev LT, Markovska IG, Lyubchev LA. Non-isothermal kinetics of pyrolysis of rice husk. Thermochimica Acta. 2003;406:1-7. DOI: 10.1016/S0040-6031(03)00222-3
  12. 12. Uslu A, Faaij APC, Bergman PCA. Pre-treatment technologies, and their effect on international bioenergy supply chain logistics. Techno-economic evaluation of torrefaction, fast pyrolysis and pelletisation. Energy. 2008;33:1206-1223. DOI: 10.1016/j.energy.2008.03.007
  13. 13. Li H, Liu X, Legros R, Bi XT, Lim CJ, Sokhansanj S. Torrefaction of sawdust in a fluidized bed reactor. Bioresource Technology. 2012;103:453-458. DOI: 10.1016/j.biortech.2011.10.009
  14. 14. Peng JH, Bi HT, Sokhansanj S, Lim JC. A study of particle size effect on biomass torrefaction and densification. Energy and Fuels. 2012;26:3826-3839. DOI: 10.1021/ef3004027
  15. 15. Bridgwater AV, Peacocke GVC. Fast pyrolysis processes for biomass. Renewable and Sustainable Energy Reviews. 2000;4:1-73. DOI: 10.1016/S1364-0321(99)00007-6
  16. 16. Bridgwater AV. Review of fast pyrolysis of biomass and product upgrading. Biomass and Bioenergy. 2012;38:68-94. DOI: 10.1016/j.biombioe.2011.01.048
  17. 17. Isahak WNRW, Hisham MWM, Yarmo MA, Yun Hin T-Y. A review on bio-oil production from biomass by using pyrolysis method. Renewable and Sustainable Energy Reviews. 2012;16:5910-5923. DOI: 10.1016/j.rser.2012.05.039
  18. 18. Angın D. Effect of pyrolysis temperature and heating rate on biochar obtained from pyrolysis of safflower seed press cake. Bioresource Technology. 2013;128:593-597. DOI: 10.1016/j.biortech.2012.10.150
  19. 19. Altun NE, Hicyilmaz C, Kök MV. Effect of particle size and heating rate on the pyrolysis of silopi asphaltite. Journal of Analytical and Applied Pyrolysis. 2003;67:369-379. DOI: 10.1016/S0165-2370(02)00075-X
  20. 20. Antal MJ, Allen SG, Dai X, Shimizu B, Tam MS, Grønli M. Attainment of the theoretical yield of carbon from biomass. Industrial and Engineering Chemistry Research. 2000;39:4024-4031. DOI: 10.1021/ie000511u
  21. 21. Jarvis JM, McKenna AM, Hilten RN, Das KC, Rodgers RP, Marshall AG. Characterization of pine pellet and peanut hull pyrolysis bio-oils by negative-ion electrospray ionization fourier transform ion cyclotron resonance mass spectrometry. Energy and Fuels. 2012;26:3810-3815. DOI: 10.1021/ef300385f
  22. 22. NikAzar M, Hajaligol MR, Sohrabi M, Dabir B. Effects of heating rate and particle size on the products yields from rapid pyrolysis of beech-wood. Fuel Science and Technology International. 1996;14:479-502
  23. 23. Lee J. Biological conversion of lignocellulosic biomass to ethanol. Journal of Biotechnology. 1997;56:1-24. DOI: 10.1016/S0168-1656(97)00073-4
  24. 24. Jenkins BM, Kayhanian M, Baxter LL, Salour D. Combustion of residual biosolids from a high solids anaerobic digestion/aerobic composting process. Biomass and Bioenergy. 1997;12:367-381. DOI: 10.1016/S0961-9534(97)00086-X
  25. 25. Radlein DSTAG, Piskorz J, Scott DS. Lignin derived oils from the fast pyrolysis of poplar wood. Journal of Analytical and Applied Pyrolysis. 1987;12:51-59. DOI: 10.1016/0165-2370(87)80014-1
  26. 26. Basu P, Rao S, Dhungana A. An investigation into the effect of biomass particle size on its torrefaction. The Canadian Journal of Chemical Engineering. 2013;91:466-474. DOI: 10.1002/cjce.21710
  27. 27. Thunman H, Leckner B, Niklasson F, Johnsson F. Combustion of wood particles—A particle model for eulerian calculations. Combustion and Flame. 2002;129:30-46. DOI: 10.1016/S0010-2180(01)00371-6
  28. 28. Scott DS, Piskorz J, Radlein D. Liquid products from the continuous flash pyrolysis of biomass. Industrial and Engineering Chemistry Process Design and Development. 1985;24:581-588. DOI: 10.1021/i200030a011
  29. 29. Demirbas A. Effects of temperature and particle size on bio-char yield from pyrolysis of agricultural residues. Journal of Analytical and Applied Pyrolysis. 2004;72:243-248. DOI: 10.1016/j.jaap.2004.07.003
  30. 30. Mani T, Murugan P, Abedi J, Mahinpey N. Pyrolysis of wheat straw in a thermogravimetric analyzer: Effect of particle size and heating rate on devolatilization and estimation of global kinetics. Chemical Engineering Research and Design: Transactions of the Institution of Chemical Engineers Part A. 2010;88:952-958. DOI: 10.1016/j.cherd.2010.02.008
  31. 31. Tsai WT, Lee MK, Chang YM. Fast pyrolysis of rice husk: Product yields and compositions. Bioresource Technology. 2007;98:22-28. DOI: 10.1016/j.biortech.2005.12.005
  32. 32. Nassar MM. Thermal analysis kinetics of bagasse and rice straw. Energy Sources. 1999;20:831-837. DOI: 10.1016/S0140-6701(99)98986-5
  33. 33. Wannapeera J, Fungtammasan B, Worasuwannarak N. Effects of temperature and holding time during torrefaction on the pyrolysis behaviors of woody biomass. Journal of Analytical and Applied Pyrolysis. 2011;92:99-105. DOI: 10.1016/j.jaap.2011.04.010
  34. 34. Tooyserkani Z, Sokhansanj S, Bi X, Lim J, Lau A, Saddler J, et al. Steam treatment of four softwood species and bark to produce torrefied wood. Applied Energy. 2013;103:514-521. DOI: 10.1016/j.apenergy.2012.10.016
  35. 35. Grønli MG, Várhegyi G, Di Blasi C. Thermogravimetric analysis and devolatilization kinetics of wood. Industrial and Engineering Chemistry Research. 2002;41:4201-4208. DOI: 10.1021/ie0201157
  36. 36. Ortega JV, Renehan AM, Liberatore MW, Herring AM. Physical and chemical characteristics of aging pyrolysis oils produced from hardwood and softwood feedstocks. Journal of Analytical and Applied Pyrolysis. 2011;91:190-198. DOI: 10.1016/j.jaap.2011.02.007
  37. 37. Heo HS, Park HJ, Park Y-K, Ryu C, Suh DJ, Suh Y-W, et al. Bio-oil production from fast pyrolysis of waste furniture sawdust in a fluidized bed. Bioresource Technology. 2010;101:S91-S96. DOI: 10.1016/j.biortech.2009.06.003
  38. 38. Lédé J, Broust F, Ndiaye F-T, Ferrer M. Properties of bio-oils produced by biomass fast pyrolysis in a cyclone reactor. Fuel. 2007;86:1800-1810. DOI: 10.1016/j.fuel.2006.12.024
  39. 39. Zhang H, Xiao R, Huang H, Xiao G. Comparison of non-catalytic and catalytic fast pyrolysis of corncob in a fluidized bed reactor. Bioresource Technology. 2009;100:1428-1434. DOI: 10.1016/j.biortech.2008.08.031
  40. 40. Cao J-P, Xiao X-B, Zhang S-Y, Zhao X-Y, Sato K, Ogawa Y, et al. Preparation and characterization of bio-oils from internally circulating fluidized-bed pyrolyses of municipal, livestock, and wood waste. Bioresource Technology. 2011;102:2009-2015. DOI: 10.1016/j.biortech.2010.09.057
  41. 41. Thangalazhy-Gopakumar S, Adhikari S, Ravindran H, Gupta RB, Fasina O, Tu M, et al. Physiochemical properties of bio-oil produced at various temperatures from pine wood using an Auger reactor. Bioresource Technology. 2010;101:8389-8395. DOI: 10.1016/j.biortech.2010.05.040
  42. 42. Heo HS, Park HJ, Dong J-I, Park SH, Kim S, Suh DJ, et al. Fast pyrolysis of rice husk under different reaction conditions. Journal of Industrial and Engineering Chemistry. 2010;16:27-31. DOI: 10.1016/j.jiec.2010.01.026
  43. 43. Mullen CA, Boateng AA, Goldberg NM, Lima IM, Laird DA, Hicks KB. Bio-oil and bio-char production from corn cobs and stover by fast pyrolysis. Biomass and Bioenergy. 2010;34:67-74. DOI: 10.1016/j.biombioe.2009.09.012
  44. 44. Islam MR, Parveen M, Haniu H. Properties of sugarcane waste-derived bio-oils obtained by fixed-bed fire-tube heating pyrolysis. Bioresource Technology. 2010;101:4162-4168. DOI: 10.1016/j.biortech.2009.12.137
  45. 45. Jensen PA, Frandsen FJ, Hansen J, Dam-Johansen K, Henriksen N, Hörlyck S. SEM investigation of superheater deposits from biomass-fired boilers. Energy and Fuels. 2004;18:378-384. DOI: 10.1021/ef030097l
  46. 46. Raveendran K, Ganesh A, Khilar KC. Influence of mineral matter on biomass pyrolysis characteristics. Fuel. 1995;74:1812-1822. DOI: 10.1016/0016-2361(95)80013-8
  47. 47. Chiaramonti D, Oasmaa A, Solantausta Y. Power generation using fast pyrolysis liquids from biomass. Renewable and Sustainable Energy Reviews. 2007;11:1056-1086. DOI: 10.1016/j.rser.2005.07.008
  48. 48. Jensen PA, Sander B, Dam-Johansen K. Pretreatment of straw for power production by pyrolysis and char wash. Biomass and Bioenergy. 2001;20:431-446. DOI: 10.1016/S0961-9534(01)00005-8
  49. 49. Oasmaa A, Peacocke C, Gust S, Meier D, McLellan R. Norms and standards for pyrolysis liquids. End-user requirements and specifications. Energy and Fuels. 2005;19:2155-2163. DOI: 10.1021/ef040094o
  50. 50. Oasmaa A, Sipilä K, Solantausta Y, Kuoppala E. Quality improvement of pyrolysis liquid: Effect of light volatiles on the stability of pyrolysis liquids. Energy and Fuels. 2005;19:2556-2561. DOI: 10.1021/ef0400924
  51. 51. Agblevor FA, Besler S. Inorganic compounds in biomass feedstocks. 1. Effect on the quality of fast pyrolysis oils. Energy and Fuels. 1996;10:293-298. DOI: 10.1021/ef950202u
  52. 52. Meier D, van de Beld B, Bridgwater AV, Elliott DC, Oasmaa A, Preto F. State-of-the-art of fast pyrolysis in Iea bioenergy member countries. Renewable and Sustainable Energy Reviews. 2013;20:619-641. DOI: 10.1016/j.rser.2012.11.061
  53. 53. Doug E. Advancement of bio oil utilization for refinery feedstock. In: Washington Bioenergy Research Symposium. 2010
  54. 54. Sipilä K, Kuoppala E, Fagernäs L, Oasmaa A. Characterization of biomass-based flash pyrolysis oils. Biomass and Bioenergy. 1998;14:103-113. DOI: 10.1016/S0961-9534(97)10024-1
  55. 55. Zhang H, Xiao R, Wang D, He G, Shao S, Zhang J, et al. Biomass fast pyrolysis in a fluidized bed reactor under N2, CO2, CO, Ch4 and H2 atmospheres. Bioresource Technology. 2011;102:4258-4264. DOI: 10.1016/j.biortech.2010.12.075
  56. 56. Brewer CE, Schmidt-Rohr K, Satrio JA, Brown RC. Characterization of biochar from fast pyrolysis and gasification systems. Environmental Progress and Sustainable Energy. 2009;28:386-396. DOI: 10.1002/ep.10378
  57. 57. Fouladi Tajar A, Kaghazchi T, Soleimani M. Adsorption of cadmium from aqueous solutions on sulfurized activated carbon prepared from nut shells. Journal of Hazardous Materials. 2009;165:1159-1164. DOI: 10.1016/j.jhazmat.2008.10.131
  58. 58. Demirbas A. Agricultural based activated carbons for the removal of dyes from aqueous solutions: A review. Journal of Hazardous Materials. 2009;167:1-9. DOI: 10.1016/j.jhazmat.2008.12.114
  59. 59. Yang Y, Lin X, Wei B, Zhao Y, Wang J. Evaluation of adsorption potential of bamboo biochar for metal-complex dye: Equilibrium, kinetics and artificial neural network modeling. International Journal of Environmental Science and Technology. 2014;11:1093-1100. DOI: 10.1007/s13762-013-0306-0
  60. 60. Park J, Hung I, Gan Z, Rojas OJ, Lim KH, Park S. Activated carbon from biochar: Influence of its physicochemical properties on the sorption characteristics of phenanthrene. Bioresource Technology. 2013;149:383-389. DOI: 10.1016/j.biortech.2013.09.085
  61. 61. Fang Q, Chen B, Lin Y, Guan Y. Aromatic and hydrophobic surfaces of wood-derived biochar enhance perchlorate adsorption via hydrogen bonding to oxygen-containing organic groups. Environmental Science and Technology. 2013;48:279-288. DOI: 10.1021/es403711y
  62. 62. Tannous K, Lam PS, Sokhansanj S, Grace JR. Physical properties for flow characterization of ground biomass from Douglas Fir wood. Particulate Science and Technology. 2012;31:291-300. DOI: 10.1080/02726351.2012.732676
  63. 63. Vamvuka D. Bio-oil, solid and gaseous biofuels from biomass pyrolysis processes—An Overview. International Journal of Energy Research. 2011;35:835-862. DOI: 10.1002/er.1804
  64. 64. Pang S, Mujumdar AS. Drying of woody biomass for bioenergy: Drying technologies and optimization for an integrated bioenergy plant. Drying Technology. 2010;28:690-701. DOI: 10.1080/07373931003799236
  65. 65. Ball RC, McIntosh A, Brindley J. The role of char-forming processes in the thermal decomposition of cellulose. Physical Chemistry Chemical Physics. 1999;1:5035-5043
  66. 66. Di Blasi C. Modeling intra- and extra-particle processes of wood fast pyrolysis. AIChE Journal. 2002;48:2386-2397. DOI: 10.1002/aic.690481028
  67. 67. Di Blasi C. Modeling chemical and physical processes of wood and biomass pyrolysis. Progress in Energy and Combustion Science. 2008;34:47-90. DOI: 10.1016/j.pecs.2006.12.001
  68. 68. Rezaei H. Physical and thermal characterization of ground wood chip and ground wood pellet particles [PhD dissertation]. Chemical and Biological Engineering, University of British Columbia (UBC); 2017
  69. 69. Gómez-de la Cruz FJ, Cruz-Peragón F, Casanova-Peláez PJ, Palomar-Carnicero JM. A vital stage in the large-scale production of biofuels from spent coffee grounds: The drying kinetics. Fuel Processing Technology. 2015;130:188-196. DOI: 10.1016/j.fuproc.2014.10.012
  70. 70. Berberović A, Milota MR. Impact of wood variability on the drying rate at different moisture content levels. Forest Products Journal. 2011;61:435-442
  71. 71. Wan Nadhari WNA, Hashim R, Danish M, Sulaiman O, Hiziroglu S. A model of drying kinetics of acacia mangium wood at different temperatures. Drying Technology. 2014;32:361-370. DOI: 10.1080/07373937.2013.829855
  72. 72. Sigge GO, Hansmann CF, Joubert E. Effect of temperature and relative humidity on the drying rates and drying times of green bell peppers (Capsicum Annuuml). Drying Technology. 1998;16:1703-1714. DOI: 10.1080/07373939808917487
  73. 73. Tapia-Blácido D, Sobral PJ, Menegalli FC. Effects of drying temperature and relative humidity on the mechanical properties of amaranth flour films plasticized with glycerol. Brazilian Journal of Chemical Engineering. 2005;22:249-256
  74. 74. Chen D, Zhang Y, Zhu X. Drying kinetics of rice straw under isothermal and nonisothermal conditions: A comparative study by thermogravimetric analysis. Energy and Fuels. 2012;26:4189-4194. DOI: 10.1021/ef300424n
  75. 75. Rezaei H, Sokhansanj S, Lim CJ, Lau A, Bi X. Effects of the mass and volume shrinkage of ground chip and pellet particles on drying rates. Particuology. 2018;38:1-9. DOI: 10.1016/j.partic.2017.09.001
  76. 76. Mujumdar AS. Handbook of Industrial Drying. Boca Raton: Tylor & Francis; 2006
  77. 77. Mazzanti P, Togni M, Uzielli L. Drying shrinkage and mechanical properties of poplar wood (Populus alba L.) across the grain. Journal of Cultural Heritage. 2012;13:S85-S89. DOI: 10.1016/j.culher.2012.03.015
  78. 78. Taylor A, Plank B, Standfest G, Petutschnigg A. Beech wood shrinkage observed at the microscale by a time series of X-Ray computed tomographs (Μxct). De Gruyter. 2013;67:201-205
  79. 79. Rezaei H, Yazdanpanah F, Lim CJ, Lau A, Sokhansanj S. Pyrolysis of ground pine chip and ground pellet particles. The Canadian Journal of Chemical Engineering. 2016;94:1863-1871. DOI: 10.1002/cjce.22574
  80. 80. Rezaei H, Sokhansanj S, Bi X, Lim CJ, Lau A. A numerical and experimental study on fast pyrolysis of single woody biomass particles. Applied Energy. 2017;198:320-331. DOI: 10.1016/j.apenergy.2016.11.032
  81. 81. Rezaei H, Lim CJ, Lau A, Sokhansanj S. Size, shape and flow characterization of ground wood chip and ground wood pellet particles. Powder Technology. 2016;301:737-746. DOI: 10.1016/j.powtec.2016.07.016
  82. 82. van der Stelt MJC, Gerhauser H, Kiel JHA, Ptasinski KJ. Biomass upgrading by torrefaction for the production of biofuels: A review. Biomass and Bioenergy. 2011;35:3748-3762. DOI: 10.1016/j.biombioe.2011.06.023
  83. 83. Li S, Xu S, Liu S, Yang C, Lu Q. Fast pyrolysis of biomass in free-fall reactor for hydrogen-rich gas. Fuel Processing Technology. 2004;85:1201-1211. DOI: 10.1016/j.fuproc.2003.11.043
  84. 84. Şensöz S, Angın D, Yorgun S. Influence of particle size on the pyrolysis of rapeseed (Brassica Napus L.): Fuel properties of bio-oil. Biomass and Bioenergy. 2000;19:271-279. DOI: 10.1016/S0961-9534(00)00041-6
  85. 85. Şensöz S, Can M. Pyrolysis of pine (Pinus Brutia Ten.) chips: 1. Effect of pyrolysis temperature and heating rate on the product yields. Energy Sources. 2002;24:347-355. DOI: 10.1080/00908310252888727
  86. 86. Encinar JM, Beltrán FJ, Bernalte A, Ramiro A, González JF. Pyrolysis of two agricultural residues: olive and grape bagasse. Influence of particle size and temperature. Biomass and Bioenergy. 1996;11:397-409. DOI: 10.1016/S0961-9534(96)00029-3
  87. 87. Scott DS, Piskorz J. The continuous flash pyrolysis of biomass. Canadian Journal of Chemical Engineering. 1984;62:404-412
  88. 88. Czernik S, Bridgwater AV. Overview of applications of biomass fast pyrolysis oil. Energy and Fuels. 2004;18:590-598. DOI: 10.1021/ef034067u
  89. 89. Meng J, Park J, Tilotta D, Park S. The effect of torrefaction on the chemistry of fast-pyrolysis bio-oil. Bioresource Technology. 2012;111:439-446. DOI: 10.1016/j.biortech.2012.01.159
  90. 90. Jendoubi N, Broust F, Commandre JM, Mauviel G, Sardin M, Lédé J. Inorganics distribution in bio oils and char produced by biomass fast pyrolysis: The key role of aerosols. Journal of Analytical and Applied Pyrolysis. 2011;92:59-67. DOI: 10.1016/j.jaap.2011.04.007
  91. 91. Scott DS, Paterson L, Piskorz J, Radlein D. Pretreatment of poplar wood for fast pyrolysis: Rate of cation removal. Journal of Analytical and Applied Pyrolysis. 2001;57:169-176. DOI: 10.1016/S0165-2370(00)00108-X
  92. 92. Piskorz J, Radlein DSAG, Scott DS, Czernik S. Pretreatment of wood and cellulose for production of sugars by fast pyrolysis. Journal of Analytical and Applied Pyrolysis. 1989;16:127-142. DOI: 10.1016/0165-2370(89)85012-0
  93. 93. Chaiwat W, Hasegawa I, Kori J, Mae K. Examination of degree of cross-linking for cellulose precursors pretreated with acid/hot water at low temperature. Industrial and Engineering Chemistry Research. 2008;47:5948-5956. DOI: 10.1021/ie800080u
  94. 94. Alvira P, Tomás-Pejó E, Ballesteros M, Negro MJ. Pretreatment technologies for an efficient bioethanol production process based on enzymatic hydrolysis: A review. Bioresource Technology. 2010;101:4851-4861. DOI: 10.1016/j.biortech.2009.11.093
  95. 95. Das P, Ganesh A, Wangikar P. Influence of pretreatment for deashing of sugarcane bagasse on pyrolysis products. Biomass and Bioenergy. 2004;27:445-457. DOI: 10.1016/j.biombioe.2004.04.002
  96. 96. Eom I-Y, Kim K-H, Kim J-Y, Lee S-M, Yeo H-M, Choi I-G, et al. Characterization of primary thermal degradation features of lignocellulosic biomass after removal of inorganic metals by diverse solvents. Bioresource Technology. 2011;102:3437-3444. DOI: 10.1016/j.biortech.2010.10.056
  97. 97. Bakker RR, Jenkins BM. Feasibility of collecting naturally leached rice straw for thermal conversion. Biomass and Bioenergy. 2003;25:597-614. DOI: 10.1016/S0961-9534(03)00053-9
  98. 98. Phanphanich M, Mani S. Impact of torrefaction on the grindability and fuel characteristics of forest biomass. Bioresource Technology. 2011;102:1246-1253. DOI: 10.1016/j.biortech.2010.08.028
  99. 99. Boateng AA, Mullen CA. Fast pyrolysis of biomass thermally pretreated by torrefaction. Journal of Analytical and Applied Pyrolysis. 2013;100:95-102. DOI: 10.1016/j.jaap.2012.12.002
  100. 100. Liaw S-S, Zhou S, Wu H, Garcia-Perez M. Effect of pretreatment temperature on the yield and properties of bio-oils obtained from the auger pyrolysis of Douglas Fir wood. Fuel. 2013;103:672-682. DOI: 10.1016/j.fuel.2012.08.016
  101. 101. Ren S, Lei H, Wang L, Bu Q, Chen S, Wu J, et al. The effects of torrefaction on compositions of bio-oil and syngas from biomass pyrolysis by microwave heating. Bioresource Technology. 2013;135:659-664. DOI: 10.1016/j.biortech.2012.06.091
  102. 102. Tumuluru JS, Sokhansanj S, Wright CT, Boardman RD. Biomass Torrefaction Process Review and Moving Bed Torrefaction System Model Development. US: Department of Energy, O.o.B.P., editor. Idaho National Laboratory; 2010
  103. 103. Liu Q, Zhong Z, Wang S, Luo Z. Interactions of biomass components during pyrolysis: A Tg-Ftir study. Journal of Analytical and Applied Pyrolysis. 2011;90:213-218. DOI: 10.1016/j.jaap.2010.12.009
  104. 104. Zheng AQ, Zhao ZL, Chang S, Huang Z, He F, Li HB. Effect of torrefaction temperature on product distribution from two-staged pyrolysis of biomass. Energy and Fuels. 2012;26:2968-2974. DOI: 10.1021/ef201872y
  105. 105. Klinger JL. Production of biofuel intermediates from woody feedstocks via fast pyrolysis and torrefaction [master of science dissertation]. Chemical Engineering, Michigan Technology University; 2011
  106. 106. Prins MJ, Ptasinski KJ, Janssen FJJG. More efficient biomass gasification via torrefaction. Energy. 2006;31:3458-3470. DOI: 10.1016/j.energy.2006.03.008
  107. 107. Ren S, Lei H, Wang L, Bu Q, Chen S, Wu J, et al. Biofuel production and kinetics analysis for microwave pyrolysis of douglas fir sawdust pellet. Journal of Analytical and Applied Pyrolysis. 2012;94:163-169. DOI: 10.1016/j.jaap.2011.12.004

Written By

Hamid Rezaei, Fahimeh Yazdanpanah, Jim Choon Lim, Anthony Lau and Shahab Sokhansanj

Submitted: 22 September 2018 Reviewed: 02 October 2018 Published: 24 December 2018