Sei sulla pagina 1di 10

Energy 139 (2017) 818e827

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

Study of reactivity reduction in sugarcane bagasse as consequence of a


torrefaction process
D.A. Granados a, b, R.A. Ruiz a, L.Y. Vega a, F. Chejne a, *
a
Universidad Nacional de Colombia, Facultad de Minas, Cr. 80 No. 65-223, Medellín, CP 050034, Colombia
b lica de Oriente, Cr. 46 No. 40B 50, Rionegro, CP 054040, Colombia
Universidad Cato

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

Article history: A complete study of torrefaction process to sugarcane bagasse under different thermal conditions was
Received 5 April 2017 developed. After torrefaction process, chars were burned in an oxidizing atmosphere and its reactivities
Received in revised form were studied. The torrefied biomass was characterized through chemical and physical analysis such as
25 July 2017
proximate and elemental analyses, lignocellulosic composition, High Heating Value (HHV), Fourier
Accepted 4 August 2017
Available online 5 August 2017
Transform Infrared Spectroscopy (FTIR) and observations in Scanning Electron Microscopy (SEM). The
torrefied biomass was also evaluated in oxidizing conditions in order to analyze the impact of torre-
faction process over biomass combustion process. Remarkable changes of the main functional groups
Keywords:
Biomass torrefaction
were observed as the severity of the torrefaction process increased. The main structural carbohydrates
Sugarcane bagasse affected in the process were hemicellulose and cellulose, which break down largely as a result of
Combustion reactivity decarboxylation reactions, and breakage of links with methyl and acetyl groups. This was evidenced
SEM qualitatively when the material was observed in SEM, an increase in the decomposition of the cell wall
FTIR structures as the process temperature increased was noted. These changes in the material were also
verified with FTIR tests, and lignified matrix with higher content of aromatic groups and greater thermal
stability is yielded with temperature increases.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction heating rates [3,5,6], is a good alternative to give residual biomass a


better use with energy purposes. Multiple technical and scientific
Residual biomass in Colombia comes from agricultural holding, reports reveal that torrefaction improves biomass properties such
such as rice husks, banana rachis, coffee wastes, sugarcane bagasse, as hydrophobicity, milling properties, increases energy density
fiber, and palm oil shell. According to energy content and avail- reducing H/C and O/C ratios, and it acquires properties very similar
ability, the sugarcane bagasse has one of the highest HHV with to low rank coals, peats and lignites [7e13]. Because of these new
around 18 MJ/kg [1]. Therefore, there is a growing interest in properties in the solid, the torrefied biomass is seen as an alter-
bagasse as a precursor for energetic purposes and there is a need to native solid fuel with possible use in other thermochemical pro-
understanding their chemical and physical properties to optimizing cesses such as combustion, pyrolysis, or gasification according to
its use in the energy sector. Biomass has some drawbacks such as previous studies [10,14e17].
high moisture content, low energy content, hygroscopic nature, low Among the wide information on biomass torrefaction processes,
density, and heterogeneous properties, resulting in low conversion it stands out the research carried out by Wilk et al. [18], where
efficiency and difficulty in the collection, crushing, storage and wood and torrefied wood were characterized by instrumental
transport [2,3]. Torrefaction, meanwhile, appears as a good alter- techniques such as SEM, FTIR, and morphological changes of
native for thermal pre-treatment because attacks the big draw- biomass and microstructure were studied. These analyses were
backs raised previously [4]. supplemented with TGA analysis, elemental and proximate char-
Torrefaction, as a thermochemical treatment driven in an inert acterization. A primary characterization revealed changes in the
atmosphere at temperatures between 200 and 300  C and low properties of the biomass, which tended to be close to those for
coal. It was also found that the moisture content decreased and the
mechanical properties were improved. On the other hand, authors
* Corresponding author. like Nocquet et al. [19] worked with thermogravimetry coupled to a
E-mail address: fchejne@unal.edu.co (F. Chejne).

http://dx.doi.org/10.1016/j.energy.2017.08.013
0360-5442/© 2017 Elsevier Ltd. All rights reserved.
D.A. Granados et al. / Energy 139 (2017) 818e827 819

system for measuring gases, which managed to characterize the characterization of a particular Colombian biomass such as the
gaseous species that are released in the torrefaction of beech logs. sugarcane bagasse from jaggery production process. This bagasse
They proposed a mechanism to explain interactions between was torrefied and burned and then characterized by chemical and
biomass constituents, which relates some process variables such as physical point of view, in order to give explanations to the behavior
temperature and residence time with the changes in the material. of bagasse during both thermal process. The sugarcane bagasse was
Chen et al. [20,21] studied the impact of torrefaction on the torrefied at four different temperatures (230, 250, 270, and 290  C),
lignocellulosic structure of bamboo and willow, and characterized while physical and chemical characterizations were performed for
the transformation of hemicellulose, cellulose and lignin by raw and torrefied biomass through proximate and ultimate anal-
thermogravimetry. ysis, HHV, FTIR and char reactivity in oxidative environment. In
FTIR technique has been useful in the characterization of bio- addition, a qualitative morphological study using SEM techniques
masses to elucidate significant changes in its microstructure after allowed to contrast the information obtained by all different
thermal processes, showing how the functional groups is changing characterization techniques. From burning tests, traditional
as the process occurs. This performs the transformation of the parameter of activation energy, reaction rate and a reactivity index
chemical structure and the change in concentration of some func- were obtained.
tional groups in the material. These changes are mainly attributed
to the devolatilization, carbonization and depolymerization [22,23] 2. Methods
of polymeric chains that are organized in an amorphous way,
whose thermal resistance is very weak [24]. 2.1. Materials
Werner et al. [25], studied the thermal decomposition of seven
different hemicelluloses. The behavior of these components was Samples of a lignocellulosic residue from sugarcane industry
analyzed by thermogravimetric analysis, FTIR, Differential Scanning was used in this study. The biomass was dried at room temperature
Calorimetry (DSC) and Pyrolysis Gas Chromatography Mass Spec- with an average temperature of 23  C and 400 W/m2 of solar
trometry (Py-GC/MS). It was found that temperatures near 300  C, average radiation for 12 h. Then, it was first crushed in a conven-
lead to the breakdown of most inter and intramolecular hydrogen tional ball mill and then in a micro mill IKA MF-10 for obtain fine
links and CeC and CeO bonds, resulting in the formation of hy- particles. All material was sieved for obtaining particle sizes
drophilic extractives, carboxylic acids, alcohols, aldehydes, ethers, ranging between 0.09 and 0.075 mm, and stored in sealed bags for
and gases such as CO, CO2, and CH4. At these temperatures, the protection after and before tests.
cellular structure is substantially destroyed, the biomass loses its
fibrous nature and becomes brittle. Similarly, Chang et al. [26]
2.2. Torrefaction process
showed that the products distribution is significantly influenced
by the temperature and chemical composition of biomass. These
Torrefaction tests were performed in a thermogravimetric
products were analyzed by different analytical techniques such as
analyzer Navas Instruments TGA-2000A, where 12 samples are
FTIR, Thermogravimetric Analysis (TGA), X-ray Diffraction (XRD),
processed simultaneously. The torrefaction process was performed
and Py-GC/MS. This allowed them to validate that thermal
in two stages. Biomass drying was developed from room temper-
decomposition of hemicellulose predominates over lignin and
ature until 105  C with a heating rate of 10  C/min during 1 h. Then,
cellulose, generating a significant removal of water and organic
a second heating until the process temperature ranging on 230 and
components of low molecular weight. This suggests the existence
290  C, with the same heating rate was developed during 3 h. A
of crosslinks and carbohydrates carbonization due to the formation
nitrogen flow of 100 ml/min was introduced to the system to
of insoluble fibers acid.
ensure the required inert atmosphere in the process. The amount of
Hoi and Martincigh [27] worked on the characterization of four
sample used in each crucible was 1 g, thus, 12 g of torrefied biomass
species of sugarcane from Mauritania, including leaves, fodder, and
was obtained after each test and stored in air-tight containers
torrefied sugarcane, by TGA, FTIR, SEM techniques and proximate
before the analyses for protection. Torrefied samples where named
and ultimate analysis. They found that most components of bagasse
as TXXX, where T stands for torrefaction process and XXX stand the
are thermally stable at temperatures below 200  C, as found by
temperature used in torrefaction.
Tumuluru et al. [24]. Hemicellulose is the constituent whose
decomposition takes place from 180  C and with a high degradation
rate, it is followed by the cellulose in the range between 240 and 2.3. Elemental, proximate and torrefaction yields
350  C, and the lignin is extensively decomposed at temperatures
even above 500  C [9,28]. The product release plays an important Raw and torrefied biomass were characterized by proximate and
role in the final composition of torrefied biomass in their carbon elemental analysis, HHV, fuel ratio (FR), Mass Yield (MY), and En-
content, volatile matter and HHV, because it defines the feasibility ergy Yield (EY). Parameters FR, MY, and EY are defined as follows:
of torrefied biomass as an applicable solid fuel in other thermo-
FC
chemical processes [29,30]. Collard et al. [22] studied the mecha- FR ¼ (1)
nisms and composition of the products obtained from the VM
decomposition of the three main components of biomass (hemi-
cellulose, cellulose and lignin) in a pyrolysis process. They analyzed MT;daf
MY ¼ (2)
the mechanisms for each polymer decomposition in different MR;daf
thermal conditions.
Combustion reactivity for torrefied biomass has been studied by MT;daf xHHVT
different authors [10,16,31e35] in order to know the impact that EY ¼ (3)
MR;daf xHHVR
the torrefaction generates on the biomass. In these studies, torre-
fied biomasses are subjected to burning process, and parameters Proximate analyses for raw and torrefied biomass were carried
such as reaction rate, activation energies and other kinetic pa- out in a muffle furnace according to ASTM D1762-84 aiming to
rameters are obtained in these tests. obtaining volatile material, fixed carbon, ash, and moisture in all
The main interest of this research was to approach the samples. This analysis was carried out twice for repeatability.
820 D.A. Granados et al. / Energy 139 (2017) 818e827

Elemental analyses were measured using an Exeter Analytical with capacity to process twelve (12) samples simultaneously. Re-
Organic Elemental Analyzer Model EA1110. Tests were conducted in sults shown in Fig. 1(a) are the averages of these twelve tests
triplicate according to ASTM D5373-02. These tests determined developed to the same sample. This averages shown in this figure
carbon, hydrogen, and nitrogen amount of the material, while the have deviations lower than 5%.
oxygen content was calculated by difference. HHV was also As is observed in this figure, mass loss increases with temper-
measured in duplicate by calorimeter bomb Parr 6100 according to ature, and at the same time the mass loss rate. A previous drying
ASTM D5865-04. step was developed to avoid the free moisture effect during tor-
refaction process. The temperature which the mass loss starts in the
2.4. Lignocellulosic and SEM analysis biomass is around 180  C, which is in agreement with experimental
results of literature [8,19,42e44]. DTG for each mass loss average
The lignocellulosic analysis for lignin, cellulose and hemicellu- was calculated (Fig. 1(b)), and you can see that the mass loss rate
lose was carried out according with NREL/TP-510-42618 regulatory increases with temperature. All samples start its decomposition at
approach, which implies the ASTM E1721-01 (2009) standards for the same temperature, but depending on final process temperature,
the determination of acid insoluble material in the biomass, and the mass loss rate is different. For 230  C, the sample start its
E1758 ASTM - 01 (2007) for carbohydrate determination by High decomposition around 180  C, and a maximum mass loss rate of
Performance Liquid Chromatography (HPLC). These analyses were 0.005 s1 was reached. When the temperature increase to 290  C,
developed for raw and torrefied biomasses in order to obtain in- the mass loss starts at the same temperature, but the maximum
formation about the biomass degradation during the process. The mass loss rate is around 0.013 s1, which is 2.6 times faster.
study of morphology of raw and torrefied biomass was carried out
using a SEM JEOL JSM-5910LV with a SEI detector and a theoretical
resolution 300.000, enabling observations until 50 nm. 3.2. Mass and energy yield

2.5. FTIR analysis Mass and energy yields and HHV for torrefied samples shown in
Fig. 1(a) are shown in Table 1. In the range of temperature between
Characterizations by FTIR technique was performed in a Spec- 230 and 250  C, it can be noted that the changes for HHV in these
trometer Spectrum Two, with an MIR source and Universal ATR samples is scarce, but in the range of 270 and 290  C this behavior is
Perkin Elmer accessory with diamond crystal. With these equip- also observed. This has already been observed in the literature [3],
ment, spectra in the infrared region between 8300 cm1 and because the low temperature range is within the medium severity
350 cm1 can be obtained, but in our particular case it was set for zone and the second range is in high severity [20]. This indicates
the 4000 cm1 to 450 cm1 region. This spectrometer has a spectral that reactions occurring in each thermal range are substantially
resolution of 0.5 cm1 standard with an accuracy in their higher different.
wavelength of 0.01 cm1 to 3000 cm1. The resulting spectrum In Table 1, the effect of the temperature over MY and EY pa-
represents the sample absorption, resulting in its molecular rameters is evident. Minimum parameter values were found in
fingerprint, due to its own functional groups. several conditions. For 290  C, MY and EY were reduced until values
of 47% and 60% respectively. In soft conditions values for MY and EY
2.6. Combustion tests of 74% and 83% were found. An inverse behavior is observed for
HHV, which increased with process severity. This takes values of
Raw and torrefied material were subjected to combustion tests 18.7 MJ/kg for soft conditions and 21.5 MJ/kg for severe conditions,
in a thermogravimetric balance in order to assess its reactivities. increasing 11% and 28% in comparison with raw material. The
Approximately 10 mg of material was burned under 4 different above-mentioned argument is another plus for confirming that the
thermal conditions with heating rates of 5, 10, 15 and 20  C/min. torrefaction process increases the amount of energy available in the
Prior to combustion process, a soft heating at a temperature of biomass with temperature [45]. However, with severe conditions
105  C for around 30 min was developed in order to obtain a the mass loss becomes considerable and the final mass of carbo-
complete drying process of the sample. A second heating ramp was naceous structure is lower. Moving the process to values very close
made from this temperature to 900  C with the considered heating to 300  C involves a high energy consumption and a significant
rate. DTG curves were obtained in order to find some differences
between tested materials and to compare them. Torrefied sample
T230 was not considered for reactivity testing, because its low
decomposition during torrefaction test. For this reason, only sam-
ples T250, T270 and T290 were tested. Kinetic parameters for
combustion process were obtained following Ozawa's iso-
conversional method [36], in which a distribution of activation
energies was obtained for the entire combustion process. In addi-
tion, a reactivity index was obtained in order to measure the
sample reactivity when the mass loss was 50% according whit
previous works found in literature [37e41]. This reactivity index
was the temperature when the sample reaches this mass loss. In the
same way, reaction rate was found from obtained kinetic parame-
ters when the sample reaches 50% of mass loss.

3. Results and discussion

3.1. Torrefaction tests Fig. 1. Thermogravimetric results for torrefaction tests with temperatures ranging on
230e290  C. (a) Dimensionless mass and temperature, and (b) DTG for thermograms
Torrefaction process was developed in a thermogravimetric unit in the same tests.
D.A. Granados et al. / Energy 139 (2017) 818e827 821

Table 1 higher amounts of biomass with coal in a co-firing process.


Mass and Energy yield and HHV for all torrefied samples. Solid fuels are usually classified by their degree of carbonization
Sample MY EY HHV with the help of diagrams such as ternary, Van Krevelen or Seyler
Raw 1.00 1.00 16.79 ± 0.04
[50,51], through ratios such as O/C and H/C from their elemental
T230 0.75 ± 0.004 0.83 ± 0.01 18.69 ± 0.04 composition. The Van Krevelen diagram in Fig. 2 shows that, after
T250 0.66 ± 0.001 0.75 ± 0.01 18.97 ± 0.03 torrefaction process, a good carbonization degree in the material is
T270 0.53 ± 0.001 0.68 ± 0.01 21.40 ± 0.01 achieved, according to photographs for each torrefied sample. In
T290 0.47 ± 0.002 0.61 ± 0 .0.01 21.51 ± 0.05
addition, according to Seyler classification [51], the torrefied
bagasse at 290  C can be classified as a lignite type. In this diagram,
it is possible to appreciate that during torrefaction process, the
mass loss. In this way, torrefaction process is a precise combination biomass loses some oxygenate and hydrogenate compounds, which
of its variables, in order to obtain not significant mass loss and to densifies the carbon content. In certain cases, the torrefied biomass
retain as much energy as possible in the product. is also called ‘charcoal’ owing to its similarities to conventional coal,
shown in Van Krevelen diagram [6,7,52,53].
3.3. Proximate and elemental analysis The change in color of the biomass during the torrefaction
process is a good indicator of the severity process. As shown in
Proximate and elemental analyses of raw and torrefied bagasse Fig. 2, brown color intensifies as the process temperature increases,
are shown in Table 2. From this table can be noted that the content but at high temperature, the biomass is dark brown or black color.
of volatile material decreases significantly with an increase in fixed The color is uniform in each sample, indicating that the torrefaction
carbon. This is subjected to the severity of the process and the process was performed homogeneously throughout the sample.
chemical nature of the biomass. Bagasse, as a lignocellulosic ma- This change in color can be attributed to many factors such as the
terial type, consists mainly of organic components (hemicellulose, lignocellulosic composition, changes in the surface of the biomass
cellulose and lignin) which, when subjected to high temperatures, that could affect the absorption properties, reflection and diffrac-
decompose into volatile products such as water, hydrogen, carbon tion of light. Some formation of some chromophoric groups and
dioxide, carbon monoxide, acetic acid, formic acid, furfural, levo- displacement of some sugar molecules of low molecular weight
glucosan, methanol, formaldehyde, etc. [22,25,46e49]. Therefore, it towards the surface of the particle can affect the color in the ma-
becomes a devolatilized material, with lower final mass, but with terial after being torrefied [54e57].
densified fixed carbon. As carbon from ultimate and proximate Some authors attribute the color change in biomass to the result
analysis increases, the final biomass has a significant increase in the of non-enzymatic browning reactions caused by condensation be-
calorific value, making it an attractive material for subsequent tween carbonyl compounds and amino, or degradation of com-
thermal processes [45]. pounds with conjugated double bonds to carbonyl groups. There
Ash content slightly increased with temperature. Table 2 also are four main routes to non-enzymatic browning, although
shows the elemental composition for raw and torrefied bagasse. chemistry of these reactions is related to the Maillard reaction:
Nitrogen content, which in this case is unrepresentative, was kept Maillard reaction, oxidation of ascorbic acid, lipid peroxidation, and
constant, and does not generate concerns about the formation of caramelization at high temperatures [58e60].
nitrogen oxides (NOX) emitted to the environment. With these
results, it is verified that carbon present in the samples increases
with the process severity. The fixed carbon fraction increases from
15.98% for raw material, to 49.17% when the material is torrefied at
290  C, and the carbon from the elemental analysis increases from
42.09% to 59.27% when torrefied at same conditions. Moreover,
oxygen and hydrogen also decrease during the process to values of
25.88% and 58.49% respectively when the process is performed at
the highest temperature. This is due to the increased release of
volatile compounds with high oxygen and hydrogen content, which
has little contribution to the HHV.
FR was also evaluated and was increased from 0.19 for raw
material to 1.06 for torrefied biomass at severe conditions (290  C).
This parameter its very important when the torrefied material is
evaluated in a combustion process, because it indicates the com-
bustion type that could occur in the process. Combustions of tor-
refied biomass with low FR may generate more emissions of CO2. In
contrast, combustion of torrefied biomass with high FR, reduces
greenhouse gases emission and thus it encourages a mixing of
Fig. 2. Van krevelen diagram for raw and torrefied biomass.

Table 2
Proximate and elemental analysis for raw and torrefied bagasse (both analysis in dry basis).

Sample Proximate analysis Elemental analysis

FC Ash VM S FR H C O N

Raw 15.98 ± 0.32 2.04 ± 0.01 81.86 ± 1.01 0.12 ± 0.01 0.20 ± 0.01 5.42 ± 0.08 42.09 ± 0.21 51.50 ± 0.33 0.18 ± 0.04
T230 23.87 ± 0.41 2.55 ± 0.01 73.46 ± 0.85 0.13 ± 0.01 0.32 ± 0.01 4.84 ± 0.05 48.21 ± 0.15 46.69 ± 0.22 0.26 ± 0.02
T250 34.12 ± 0.22 3.00 ± 0.02 62.74 ± 0.42 0.14 ± 0.01 0.54 ± 0.01 3.83 ± 0.02 50.92 ± 0.41 45.00 ± 0.49 0.25 ± 0.06
T270 45.07 ± 0.30 3.88 ± 0.03 50.91 ± 0.77 0.15 ± 0.01 0.89 ± 0.02 2.76 ± 0.04 57.19 ± 0.22 39.75 ± 0.29 0.30 ± 0.03
T290 49.17 ± 0.55 4.43 ± 0.02 46.25 ± 0.62 0.16 ± 0.01 1.06 ± 0.03 2.25 ± 0.09 59.27 ± 0.34 38.17 ± 0.46 0.31 ± 0.03
822 D.A. Granados et al. / Energy 139 (2017) 818e827

3.4. Lignocellulosic analysis

Lignocellulosic composition of the structure for raw and torre-


fied material was conducted according to international standards.
With this analysis, its possible to determine the biomass decom-
position degree in function of its polymeric composition. Results
shown in Table 3 correspond to the standardization of quantities
based on these three components.
As shown in Table 3, the decomposition of the hemicellulose is
evident as the temperature increases in the process. At low tem-
peratures (230  C), the decomposition becomes considerable; it
degraded about 50% of its initial amount. At higher temperatures,
the destruction of this polymer is about 100% when the tempera-
ture reaches 290  C. In the case of cellulose, its degradation effect is
not considerable at low temperatures, only when the temperature
of 250  C is reached a strong degradation of this polymer initiates,
consistent with that found by different authors [61,62]. At low Fig. 3. IR spectra for raw and torrefied biomass at 230, 250, 270, and 290  C.
temperatures, the main component that devolatilizes is hemicel-
lulose [22,25,46e49]; while cellulose and lignin are more ther-
mostable in these conditions and are less devolatilized. In severe a lesser extent cellulose and lignin. During torrefaction these
thermal conditions higher than 250  C, considerable mass losses polymers can melt, vaporize, re-polymerize, form char or either
were obtained because, in addition to hemicellulose, cellulose takes react [66]. This causes changes in the biomass properties and
greater participation in the devolatilization [20], mainly its amor- structure depending on their composition.
phous part. The peaks in the IR spectrum correspond directly to the func-
Temperature has a very noticeable effect on the biomass com- tional groups characteristic of the sample and its magnitude is
ponents, especially hemicellulose. Bergman et al. [63], associated indicative of the formation or disappearance, transformations, or
the changes in the treated biomass to certain temperature ranges as chemical reactions of them during process. The allocation of bands
follows: (1) 100e150  C, this stage is known as a nonreactive drying in the IR spectrum for raw and torrefied bagasse was made based
where the chemical components of the biomass remain almost on a literature review where different species of bagasse were
intact and it only causes the removal of surface moisture. This re- characterized by FTIR [27,46,67e73]. A more detailed description of
duces porosity and small structural ruptures appear; (2) from 150 the FTIR peaks based on the main polymer components of the
to 200  C, it is known as a stage of reactive drying and at this point, biomass (hemicellulose, cellulose and lignin) is shown in Table 4.
lignin begins a soft decomposition. At these temperatures, struc- In Fig. 3 you can see the spectrum of raw and torrefied biomass,
tural distortions change the original structure of the biomass, by and a wide band is displayed on 3334 cm1 approximately for all
breaking hydrogen bonds and carbon, resulting in emissions of samples, and it is attributed to the vibration of OeH bonds [74]. This
extracts and lipophilic compounds; (3) 200e300  C, it is the band is caused by the presence of alcoholic, hydroxyl, and phenolic
destructive drying zone, resulting in charring and devolatilization groups involved in hydrogen bonds of the lignin and certain car-
of the sample. At these temperatures, the cell structure is largely bohydrates. Its intensity decreases by methylation reactions, but
destroyed losing its fibrous nature, so it becomes more fragile. increases with demethylation reactions [75]. This indicates that
Below 250  C, the mass loss is mainly attributed to the hemicel- methylation occurs in the sample due to torrefaction process. In
lulose. Above 250  C, the hemicellulose and cellulose decomposi- addition, the loss of this functional group in the material during the
tion occurs mainly in volatile matter. Lignin has a small torrefaction process improves hydrophobicity because of the for-
devolatilization in this condition [20e22,24,25,53,62,64,65]. mation of nonpolar and unsaturated compounds.
A small band is observed in 2917 cm1 attributed to the CeH
stretching of aliphatic groups methyl, methylene, or methane
3.5. FTIR analysis
(CHal) in the lignin. The peak at 1726 cm1 is associated to the
stretching of the links C]O conjugated and unconjugated
The FTIR technique was used to investigate the torrefaction ef-
(carbonyl/carboxyl) of carboxylic acids in hemicellulose. Combined
fect on the chemical structure of bagasse. IR spectra for raw and
with this peak, can appear structures generated by the dehydration
torrefied bagasse are shown in Fig. 3, where it is possible to observe
of cellulose (C]O), which can also be observed around
that the changes in the bands occur in the proximity of the 4000
1620 cm1 (C]C), and which increase with the severity of the
and 600 cm1, since organic components have fundamental vi-
process. Previous work [69,76] suggests that the torrefaction
bration bands in the infrared middle region of 4000 and 200 cm1.
removes this signal by decreasing the amount of acetyl groups,
The changes are mainly due to degradation of polymer components
which leads to new products that appear in low wavenumbers
of bagasse such as hemicellulose, consisting mostly of xylan, and to
(1700 cm1) as evidenced in this work. As reported by Chang et al.
[76], at values close to 1632 cm1 a peak for absorbed water should
Table 3 appear, but in our work it was not detected probably due to some
Lignocellulosic composition for raw and torrefied bagasse at 230, 250, 270, and overlap with another nearby peak.
290  C (dry and ash free basis).
Peaks located between 1600 and 1506 to 1514 cm1 are assigned
Sample Hemicellulose (wt.%) Cellulose (wt.%) Lignin (wt.%) to vibrations in aromatic skeletons of lignin (C]C) and they are
Raw 26.46 ± 1.74 28.25 ± 0.43 20.44 ± 0.15 maintained with an increasing tendency as temperature increases.
T230 13.13 ± 0.03 31.09 ± 0.94 46.20 ± 0.49 The latter occurs, firstly because lignin does not suffer significant
T250 6.39 ± 0.85 23.74 ± 0.78 62.71 ± 1.20 modifications and secondly, because cellulose and hemicellulose
T270 2.80 ± 0.14 4.97 ± 0.09 80.39 ± 1.98 degrade to form primary char rich in aromatic groups, which are
T290 1.44 ± 0.14 0.23 ± 3.22 94.51 ± 1.51
D.A. Granados et al. / Energy 139 (2017) 818e827 823

Table 4
Main band assignments of the FTIR spectra of the raw and torrefied bagasse [27,46,67e73].

Wave Number (cm1) Assignment/functional group Polymer

3400 OeH Stretching bonds Lignin


2800e2900 CeH Stretching of Methyl (eCH3) and Methylene groups (eCH2e) Lignin
1735e1739 Stretching of C]O (C]O) of oxygenated functionalities in conjugated and unconjugated Hemicellulose
systems (carbonyl/carboxyl groups)
1632 HeOeH inflections of adsorbed water, C]C Lignin
1600 and 1506 - 1510 Vibrations aromatic lignin skeletons (C]C) and vibrations of C]O Lignin
1424e1428 Deformations of CeH and carbohydrates (dCHal) Lignin
1370 Deformations of CeH Cellulose, hemicellulose, lignin
1310 Movements (Wagging) in CH2 Cellulose, hemicellulose
1280 Flection in CeH Crystalline cellulose
1271 Aromatic rings stretching (CeO) Guaiacyl lignin
1159 Asymmetrical stretching CeOeC and CeOH Cellulose, hemicellulose
1031e1034 CeO, C]C, and CeCeO Stretching Cellulose, hemicellulose, lignin
898 Deformation of CeH in amorphous cellulose xyloglucan Cellulose, hemicellulose
800e900 Glucosidic linkage Hemicellulose
664 OeH groups stretching outside of the plane Cellulose, hemicellulose

more thermostable. This leads to a greater uniformity in the the particle structure is compact and the biomass has a fibrillary
structure. organization of smooth, homogeneous parts and some rough ones
Peaks from deflections and deformations in the CeH bonds, in its external surface. As the zoom is increased, it is possible to
correspondent to some amorphous polysaccharides of lignin and notice some damage to the material at the end of some particles (a),
cellulose are observed around 1456 and 1431 cm1. Soft vibrations perhaps due to the grinding pre-treatment carried out prior to heat
can be seen in values close to 1370 cm1, which are smoothed due treatment. Increasing the zoom to 1000 and 1500, it is possible
to deformation in amorphous cellulose (CeH). Vibrations between to capture some of the internal structure, which is seen in a sheet-
1268 and 1163 cm1 are due to the stretching of the aromatic rings like form with some well-defined porous ovals (b) interconnecting
of guaiacyl lignin (CeO) and stretching in the antisymmetric oxy- in a radial direction to form a complete particle.
gen bridge CeOH and CeOeC of the cellulose and hemicellulose. In the torrefied sample at 230  C you can see, even with the
The intensity of these bands in this region tends to disappear smallest zoom, some visible damages to their external structure (c).
because of torrefaction, while a deformation of syringyl rings and a Partial warps are appreciated (d), but not very strong at this low
decomposition of xylan hemicellulose occurs. process severity. When bagasse is torrefied at 250  C, some cracks
One of the most pronounced bands in the region close to 1031 in the outer walls of the torrefied bagasse are displayed (e), and
and 1034 cm1 occurs due to vibrations in CeO, C]C and CeCeO some appearances of pores (f), possibly generated by the devola-
form groups of cellulose, hemicellulose and lignin. The intensity of tilization of less thermostable polymers. These new pores are
these peaks is stable below 230  C, but as the sample was heated apparently different from those observed in the micrographs of raw
above 250  C, the intensity of these groups decreased. Lignin is an material, since these latter are rounded.
aromatic compound in its chemical nature, so it degrades much less When those particles treated with higher thermal conditions,
than hemicellulose and cellulose, and decomposes producing i.e., 270 and 290  C, are observed, it is possible to observe much
phenols due to the breaking of the ether linkages. A very smooth more marked and deeper cracks in the torrefied particles (g, j). It is
peak near 898 cm1 is assigned to beta-glycosidic type links in also noticeable that those outer surfaces of the particles, originally
glucose from cellulose and hemicellulose. This signal decreases flat, are now presented with some deteriorated parts, large material
with process temperature. Finally, a small band at 664 cm1 detachments, and merging of some pores in these sections (h, i). A
attributed to the vibration of OeH groups out of the plane was special micrograph was obtained for a torrefied sample at 290  C,
found, but degraded as temperature increased. where the cell structure is observed with considerable damage, and
For all these peaks located below 1726 cm1, you can see that for some obvious deformations (k) which confirm the severity of the
temperatures of 230  C their intensities do not change, remaining torrefaction process all over the biomass structure.
approximately constant. The main change observed for tempera- It is noteworthy to mention that pores, cracks and crater for-
ture of 230  C was a slight decrease in the intensity of the peaks mation is a direct consequence of pronounced volatile material
located at 3344 cm1 and 2917 cm1, corresponding to light volatile release, and it is increased with process temperature. This is
molecules that are expelled at low process temperatures. Above consistent with the information found in the proximate analysis
230  C, changes in all functional groups of biomasses become (Table 2), where volatile matter content decreases as torrefaction
considerable. occurs in more severe conditions. Similarly, the cell structure ap-
pears to be melted, shrunk and almost entirely destroyed, due to
3.6. SEM analysis the degradation of hemicellulose and amorphous cellulose portion
[27].
The SEM micrographs obtained from raw and torrefied bagasse
for different thermal conditions are summarized in Fig. 4. In gen- 3.7. Combustion tests
eral, physical changes are observed in the morphology of the
samples by alterations in structure and cellular tissue. From images, Reactivity tests were performed for raw and torrefied biomass at
it was possible to identify that torrefaction leads to degradation of 250, 270 and 290  C. These tests, as described in previous sections,
the bagasse due to devolatilization, depolymerization, and were carried out with two ramps of temperature. The first ramp at
carbonization reactions of hemicellulose, cellulose and lignin [24], 105  C with 30 min of residence time, and a second ramp from
which increases when the process temperature rises. 105  C to 900  C with heating rates of 5, 10, 15, 20  C/min. In Fig. 5,
In the micrographs for raw material, it is possible to notice that the TGA and DTG for reactivity testing performed on all samples for
824 D.A. Granados et al. / Energy 139 (2017) 818e827

Zoom
100X 300X 1000X 1500X

a b
Raw

c
c d
230

e
250 e
f

g
h i
270

j k j
290

Fig. 4. SEM micrographs for raw and torrefied bagasse at 230, 250, 270, and 290  C.

a heating rate of 5  C/min are shown. The mass loss of raw material which are responsible for volatile material generation when the
starts earlier than other samples, as consequence of its untreated biomass is heated by increasing the material reactivity.
polymeric structure with high volatile matter. This structure con- For torrefied samples, the weak functional groups (e.g. CeH and
tains weak functional groups, such as CeH and OeH (see Fig. 3) OeH) were partially removed, causing the begin of mass loss later
than raw material. For the torrefied sample at 250  C, hemicellulose
and cellulose decompose partially and only 6.88 and 25.57%
respectively remain in the sample (see Table 3). This can be ratified
in Fig. 3 for sample T250, where unstable functional groups like CH,
OeH, and CeOH are present in the sample. In addition, some stable
groups formed during the torrefaction process from primary re-
actions like some aromatic compounds (shown in Fig. 3) around
700-750 cm1 and 1515-1626 cm1, contribute to decrease the
reactivity in comparison with raw biomass. The last material to
begin its mass loss was that torrefied at 290  C, because it does not
have unstable polymers such as hemicellulose and cellulose in its
structure (see Table 3), and has a lignified structure result of pri-
mary reactions and aromatic groups formed during the process.
In Fig. 5, the DTG for raw biomass shows a large initial peak
around 221  C, because of the large amount of hemicellulose in its
structure compared to torrefied samples. This peak decrease in the
other samples, indicating that the amount of hemicellulose in the
torrefied materials decreased significantly, to the point of almost
fading in the T290 sample. This behavior is consistent with shown
in Table 3. An unexpected behavior was presented in this curve at
about 264  C because a small additional peak was observed in the
Fig. 5. TGA and DTG for raw and torrefied samples for a heating rate of 5  C/min. Blue torrefaction process. This could be attributed to the amorphous
line: sample T290; green line: sample T270; red line: sample T230; black line: raw
structure of the cellulose, less thermally stable than the crystalline
sample. (For interpretation of the references to colour in this figure legend, the reader
is referred to the web version of this article.)
cellulose, which begins its decomposition at this lower
D.A. Granados et al. / Energy 139 (2017) 818e827 825

increases considerably, which is clearly observed in the peaks of the


torrefied materials at higher temperatures. In addition, this is
consistent with the aforementioned and claimed by different au-
thors [78,79], who found that the hemicellulose and cellulose
polymers, because of their primary decomposition, yield a char
primarily formed by aromatic structures that provide and increase
the aromatic structures in the material.
This behavior in the peak separation for each of the polymers in
the material is possible to be observed at low heating rates because
at high heating rates this behavior was avoided, the peaks of the
cellulose and hemicellulose tend to overlap, such as it is shown in
Fig. 6. In this figure, the sample T290 with a heating rate of 20  C/
min, only shows a broad peak where the decomposition of each
component in the process is impossible to differentiate.
The results presented in this study of reactivity can be compared
with those obtained in the FTIR analysis, where increases are
observed in the number of those peaks that reflect aromatic com-
pounds, evidenced by peaks 1600 cm1, 1506-1510 cm1, and
Fig. 6. DTG for raw and torrefied samples for all heating rates. Blue line: 20  C/min;
1271 cm1. These peaks clearly rise with increases in the temper-
green line: 15  C/min; red line 10  C/min; black line: 5  C/min. (For interpretation of
the references to colour in this figure legend, the reader is referred to the web version ature of the process, and evidently, a decrease is observed for
of this article.) groups of those less thermally stable polymers as the ones present
at 3400 cm1 and 2800-2900 cm1. These results shown in Fig. 3,
match with the obtained in these reactivity testing, volatiles
temperature and can be observed separated at a peak due to the directly reflect the reactivity of the material, but when the
low heating rate at which the process was conducted, in accordance aromaticity of the material increases the reactivity decreases.
with results from Jiang et al. [77]. At higher heating rates, this small Activation energies that describe the behavior of each material
peak was not observed. in a combustion process were obtained and are shown in Fig. 7(a).
A peak presented at temperature around 303  C corresponds to The procedure was performed by the Ozawa iso-conversional
the breakdown of cellulose. In the untreated material sample, this method [36], by means of which a distribution of activation en-
peak is shown as larger because, as it was previously mentioned ergies for the entire spectrum of material conversion can be ob-
and shown in Table 3, the amount of cellulose material is 100% of tained in each thermal condition of the tests. Kinetic parameters
the original amount. An interesting behavior in this cellulose peak are shown in Table 5 for a conversion level of 50% in the process and
is that the treatment does not generate great impact therein at a reaction order one.
temperatures below 250  C, as the peak for the T250 material is just Although many authors measure the materials reactivity when
below of that for raw material. When the material is treated at the maximum mass loss rate is reached in the process, in this work
higher temperatures, i.e., samples T270 and T290, the amount of it was chosen to measure it at 50% of the conversion, according with
cellulose is affected, as it is evidenced at the peak of these two experimental work from different authors [37e41]. In Fig. 7(b), the
samples, which decreases considerably. variation of reactivity with the pre-treatment can be observed. This
The final peak in DTG curves is attributed to lignin and other tendency was consistent with the above discussion, where the
carbonaceous materials produced in primary reactions of cellulose reactivity of the treated material decreases with the torrefaction
and hemicellulose during torrefaction. In the untreated material, temperature due to the lower presence of thermally unstable
this peak is lower than those presented in the other samples. Due to compounds after the process and more energy requirements for
decomposition of the less thermally stable polymers such as cel- obtaining a conversion level in comparison with other samples. The
lulose and hemicellulose, the proportion of lignin in the material reactivity index indicated in Fig. 7(b) corresponds to a temperature

(a) (b)
Fig. 7. Activation energy distribution from Ozawa iso-conversional method (a), and reaction rate and reactivity index (b), for raw and torrefied bagasse in a combustion process.
826 D.A. Granados et al. / Energy 139 (2017) 818e827

Table 5 - introduction. Recent adv. Thermo-chemical convers. Biomass. first ed., vol. 1.
Kinetic parameters for torrefied sugarcane bagasse. Elsevier B.V.; 2015. p. 3e29. http://dx.doi.org/10.1016/B978-0-444-63289-
0.00015-6.
Sample Ea (kJ/kmol) A (s1) [5] Chen W-H. Torrefaction. Elsevier B.V.; 2015. http://dx.doi.org/10.1016/B978-
0-12-800080-9.00010-4.
Raw 171,00 1,05Eþ17
[6] Chew JJ, Doshi V. Recent advances in biomass pretreatment e torrefaction
T250 218,55 8,35Eþ17
fundamentals and technology. Renew Sustain Energy Rev 2011;15:4212e22.
T270 166,29 3,42Eþ12 http://dx.doi.org/10.1016/j.rser.2011.09.017.
T290 143,53 2,80Eþ10 [7] Batidzirai B, Mignot PR, Schakel WB, Junginger HM, Faaij PC. Biomass torre-
faction technology: techno-economic status and future prospects. Energy
2013;62:196e214. http://dx.doi.org/10.1016/j.energy.2013.09.035.
[8] Prins MJ, Ptasinski KJ, Janssen FJJG. Torrefaction of wood. Part 1. Weight loss
value where the decomposition reaches the value of 50%. While the kinetics. J Anal Appl Pyrolysis 2006;77:28e34. http://dx.doi.org/10.1016/
reactivity index is higher, the evaluated material presents a lower j.jaap.2006.01.002.
[9] Acharya B, Sule I, Dutta A. A review on advances of torrefaction technologies
reactivity. This result was expected according to the results pre-
for biomass processing. Biomass Convers Biorefinery 2012:349e69. http://
sented in the lignocellulosic analysis, and in the results of the FTIR dx.doi.org/10.1007/s13399-012-0058-y.
analyses of the torrefied biomass. [10] Fisher EM, Dupont C, Darvell LI, Commandre  J-M, Saddawi A, Jones JM, et al.
Combustion and gasification characteristics of chars from raw and torrefied
biomass. Bioresour Technol 2012;119:157e65. http://dx.doi.org/10.1016/
4. Conclusions j.biortech.2012.05.109.
[11] Li MF, Chen CZ, Li X, Shen Y, Bian J, Sun RC. Torrefaction of bamboo under
nitrogen atmosphere: influence of temperature and time on the structure and
An important Colombian agricultural residue such as sugarcane
properties of the solid product. Fuel 2015;161:193e6. http://dx.doi.org/
bagasse was subjected to torrefaction and then to a combustion 10.1016/j.fuel.2015.08.052.
process. Kinetic parameters that give information about the reac- [12] Kastanaki E, Vamvuka D. A comparative reactivity and kinetic study on the
combustion of coalebiomass char blends. Fuel 2006;85:1186e93. http://
tivity or thermal behavior of the raw and torrefied sugarcane
dx.doi.org/10.1016/j.fuel.2005.11.004.
bagasse in combustion process were obtained. A complete physical [13] Min F, Zhang M, Zhang Y, Cao Y, Pan WP. An experimental investigation into
and chemical characterization of torrefied sugarcane bagasse was the gasification reactivity and structure of agricultural waste chars. J Anal Appl
performed to explain the decrease in reactivity when subjected to a Pyrolysis 2011;92:250e7. http://dx.doi.org/10.1016/j.jaap.2011.06.005.
[14] Chen WH, Chen CJ, Hung CI, Shen CH, Hsu HW. A comparison of gasification
combustion process. phenomena among raw biomass, torrefied biomass and coal in an entrained-
From the analysis of torrefied sugarcane bagasse, it was flow reactor. Appl Energy 2013;112:421e30. http://dx.doi.org/10.1016/
observed that the torrefaction produces a carbon enrichment, j.apenergy.2013.01.034.
[15] Mi B, Liu Z, Hu W, Wei P, Jiang Z, Fei B. Bioresource Technology Investigating
accompanied by a decrease on the functionalities of hydrogen and pyrolysis and combustion characteristics of torrefied bamboo, torrefied wood
oxygen in the char, and increases in HHV of about 30%. However, and their blends. Bioresour Technol 2016;209:50e5. http://dx.doi.org/
the mass loss becomes considerable when temperature increases, 10.1016/j.biortech.2016.02.087.
[16] Liu Z, Hu W, Jiang Z, Mi B, Fei B. Investigating combustion behaviors of
making the MY parameter to decrease considerably. In addition, the bamboo, torrefied bamboo, coal and their respective blends by thermogra-
reactivity of raw biomass is high because of volatile material in its vimetric analysis. Renew Energy 2016;87:346e52. http://dx.doi.org/10.1016/
structure. Torrefaction severity decreases the biomass reactivity j.renene.2015.10.039.
[17] Antonio Bizzo W, Lenço PC, Carvalho DJ, Veiga JPS. The generation of residual
because of volatiles deployed in the process and char formation
biomass during the production of bio-ethanol from sugarcane, its character-
from polymers. This shows that the solid product of torrefaction ization and its use in energy production. Renew Sustain Energy Rev 2014;29:
(char) has good properties when being considered as a solid fuel 589e603. http://dx.doi.org/10.1016/j.rser.2013.08.056.
[18] Wilk M, Magdziarz A, Kalemba I. Characterisation of renewable fuels' torre-
because of the properties similar to those of coal, with great po-
faction process with different instrumental techniques. Energy 2015;87:
tential for the energy industry. 259e69. http://dx.doi.org/10.1016/j.energy.2015.04.073.
The changes in the main functional groups (OeH, C]O, C]C, [19] Nocquet T, Dupont C, Commandre JM, Grateau M, Thiery S, Salvador S. Volatile
CeH and CeOeC) shown in the FTIR spectrum, are indicative of species release during torrefaction of wood and its macromolecular constit-
uents: Part 1-Experimental study. Energy 2014;72:180e7. http://dx.doi.org/
structural changes in hemicellulose, cellulose and lignin. By 10.1016/j.energy.2014.02.061.
increasing the severity of the process, the two main structural [20] Chen W-H, Kuo P-C. A study on torrefaction of various biomass materials and
carbohydrates (hemicellulose and cellulose) have been mostly its impact on lignocellulosic structure simulated by a thermogravimetry.
Energy 2010;35:2580e6. http://dx.doi.org/10.1016/j.energy.2010.02.054.
affected in the sugarcane bagasse. Thus, the char produced has [21] Chen W-H, Kuo P-C. Torrefaction and co-torrefaction characterization of
been enriched in aromatic compounds, with the almost complete hemicellulose, cellulose and lignin as well as torrefaction of some basic con-
disappearance of aliphatic compounds, which is interpreted as an stituents in biomass. Energy 2011;36:803e11. http://dx.doi.org/10.1016/
j.energy.2010.12.036.
increase in lignified material by degradation of cellulose and [22] Collard FX, Blin J. A review on pyrolysis of biomass constituents: mechanisms
hemicellulose. and composition of the products obtained from the conversion of cellulose,
hemicelluloses and lignin. Renew Sustain Energy Rev 2014;38:594e608.
http://dx.doi.org/10.1016/j.rser.2014.06.013.
Acknowledgments [23] Joshi Y, Di Marcello M, De Jong W. Torrefaction: mechanistic study of con-
stituent transformations in herbaceous biomass. J Anal Appl Pyrolysis
D.A Granados wish to thank the Colombian Administrative 2015;115:353e61. http://dx.doi.org/10.1016/j.jaap.2015.08.014.
[24] Tumuluru JS, Sokhansanj S, Hess JR, Wright CT, Boardman RD. A review on
Department of Science, Technology and Innovation (COLCIENCIAS) biomass torrefaction process and product properties for energy applications.
(Departamento Administrativo de Ciencia, Tecnología e Innovacion) Ind Biotechnol 2011;7:384e401. http://dx.doi.org/10.1089/ind.2011.0014.
for financial support. [25] Werner K, Pommer L, Brostro €m M. Thermal decomposition of hemicelluloses.
J Anal Appl Pyrolysis 2014;110:130e7. http://dx.doi.org/10.1016/
j.jaap.2014.08.013.
References [26] Chang S, Zhao Z, Zheng A, He F, Huang Z, Li H. Characterization of products
from torrefaction of sprucewood and bagasse in an auger reactor. Energy &
[1] Unidad de Planeacio  n Minero Energe tica (UPME). Atlas del Potential Ener- Fuels 2012;26:7009e17. http://dx.doi.org/10.1021/ef301048a.
tico de la Biomasa Residual en Colombia. first ed. UPME; 2010.
ge [27] Wong Sak Hoi L. Martincigh BS. Sugar cane plant fibres: separation and
[2] van der Stelt MJC, Gerhauser H, Kiel JHA, Ptasinski KJ. Biomass upgrading by characterisation. Ind Crops Prod 2013;47:1e12. http://dx.doi.org/10.1016/
torrefaction for the production of biofuels: a review. Biomass Bioenergy 2011: j.indcrop.2013.02.017.
5. http://dx.doi.org/10.1016/j.biombioe.2011.06.023. [28] Mohan D, Pittman CU, Steele PH. Pyrolysis of wood/biomass for bio-oil: a
[3] Chen W, Peng J, Bi XT. A state-of-the-art review of biomass torrefaction, densi critical review. Energy & Fuesl 2006;20:848e89. http://dx.doi.org/10.1021/
fi cation and applications. Renew Sustain Energy Rev 2015;44:847e66. http:// ef0502397.
dx.doi.org/10.1016/j.rser.2014.12.039. [29] Basu P. Biomass gasification, pyrolysis and torrefaction. Elsevier; 2013. http://
[4] Michael Sto€ cker RSAPTB. Advances in thermochemical conversion of biomass dx.doi.org/10.1016/B978-0-12-396488-5.00004-6.
D.A. Granados et al. / Energy 139 (2017) 818e827 827

[30] Toptas A, Yildirim Y, Duman G, Yanik J. Combustion behavior of different [56] Dietenberger MA, Hasburgh LE. Wood products: thermal degradation and fire.
kinds of torrefied biomass and their blends with lignite. Bioresour Technol Elsevier Ltd; 2016. http://dx.doi.org/10.1016/B978-0-12-803581-8.03338-5.
2015;177:328e36. http://dx.doi.org/10.1016/j.biortech.2014.11.072. [57] Aydemir D, Gunduz G, Ozden S. The influence of thermal treatment on color
[31] Bridgeman TG, Jones JM, Shield I, Williams PT. Torrefaction of reed canary grass, response of wood materials. Color Res Appl 2012;37:148e53. http://
wheat straw and willow to enhance solid fuel qualities and combustion prop- dx.doi.org/10.1002/col.20655.
erties. Fuel 2008;87:844e56. http://dx.doi.org/10.1016/j.fuel.2007.05.041. [58] Willits C, Underwood J, Lento H, Ricciuti C. Browning of sugar solutions I.
[32] Costa FF, Wang G, Costa M. Combustion kinetics and particle fragmentation of Effect of pH and type of amino acid in dilute sugar solutions. J Food Sci
raw and torrified pine shells and olive stones in a drop tube furnace. Proc 1958;23:61e7.
Combust Inst 2014;35:3591e9. http://dx.doi.org/10.1016/j.proci.2014.06.024. [59] Lento H, Underwood J, Willits C. Browning of sugar solutions II. Effect of the
[33] Fang MX, Shen DK, Li YX, Yu CJ, Luo ZY, Cen KF. Kinetic study on pyrolysis and position of amino group in the acid molecule in dilute gluscose solutions.
combustion of wood under different oxygen concentrations by using TG-FTIR J Food Sci 1958;23:68e71.
analysis. J Anal Appl Pyrolysis 2006;77:22e7. http://dx.doi.org/10.1016/ [60] Underwood J, Lento H, Willits C. Browning of sugar solutions 3. Effect of pH on
j.jaap.2005.12.010. the color preduced in dilute gluscose solutions containing amino acids with
[34] Jones JM, Bridgeman TG, Darvell LI, Gudka B. Saddawi a., Williams a. Combus- the amino group in different positions in the molecule. J Food Sci 1958;24:
tion properties of torrefied willow compared with bituminous coals. Fuel Pro- 181e4.
cess Technol 2012;101:1e9. http://dx.doi.org/10.1016/j.fuproc.2012.03.010. [61] Chen W-H, Kuo P-C. Isothermal torrefaction kinetics of hemicellulose, cellu-
[35] Ramajo-Escalera B, Espina A, García JR, Sosa-Arnao JH, Nebra SA. Model-free lose, lignin and xylan using thermogravimetric analysis. Energy 2011;36:
kinetics applied to sugarcane bagasse combustion. Thermochim Acta 6451e60. http://dx.doi.org/10.1016/j.energy.2011.09.022.
2006;448:111e6. http://dx.doi.org/10.1016/j.tca.2006.07.001. [62] Stefanidis SD, Kalogiannis KG, Iliopoulou EF, Michailof CM, Pilavachi PA,
[36] Ozawa TA. New method of analyzing thermogravimetric data. Bull Chem Soc Lappas AA. A study of lignocellulosic biomass pyrolysis via the pyrolysis of
Jpn 1965;38:1881e6. http://dx.doi.org/10.1246/bcsj.38.1881. cellulose, hemicellulose and lignin. J Anal Appl Pyrolysis 2014;105:143e50.
[37] Conesa JA, Marcilla A, Caballero JA, Font R. Comments on the validity and http://dx.doi.org/10.1016/j.jaap.2013.10.013.
utility of the different methods for kinetic analysis of thermogravimetric data. [63] Bergman P, Kiel J. Torrefaction for biomass upgrading. In: 14th eur biomass
J Anal Appl Pyrolysis 2001;58e59:617e33. http://dx.doi.org/10.1016/S0165- conf exhib Paris, Fr 17-21; Oct 2005. p. 3e8.
2370(00)00130-3. [64] García R, Pizarro C, Lavín AG, Bueno JL. Biomass proximate analysis using
[38] Du S-W, Chen W-H, Lucas JA. Pretreatment of biomass by torrefaction and thermogravimetry. Bioresour Technol 2013;139:1e4. http://dx.doi.org/
carbonization for coal blend used in pulverized coal injection. Bioresour 10.1016/j.biortech.2013.03.197.
Technol 2014;161:333e9. http://dx.doi.org/10.1016/j.biortech.2014.03.090. [65] Saldarriaga JF, Aguado R, Pablos A, Amutio M, Olazar M, Bilbao J. Fast char-
[39] Ebrahimi-Kahrizsangi R, Abbasi MH. Evaluation of reliability of coats-redfern acterization of biomass fuels by thermogravimetric analysis (TGA). Fuel
method for kinetic analysis of non-isothermal TGA. Trans Nonferrous Met Soc 2015;140:744e51. http://dx.doi.org/10.1016/j.fuel.2014.10.024.
China (English Ed) 2008;vol. 18:217e21. doi:10.1016/S1003-6326(08)60039-4. [66] Tumuluru JS, Sokhansanj S, Wright CT, Kremer T. GC analysis of volatiles and
[40] Ren S, Lei H, Wang L, Bu Q, Chen S, Wu J. Thermal behaviour and kinetic study for other products from biomass torrefaction process. Adv Gas Chromatogr Agric
woody biomass torrefaction and torrefied biomass pyrolysis by TGA. Biosyst Biomed Indistrial Appl 2012:211e34. http://dx.doi.org/10.5772/33488.
Eng 2013;116:420e6. http://dx.doi.org/10.1016/j.biosystemseng.2013.10.003. [67] Pohlmann JG, Oso rio E, Vilela ACF, Diez MA, Borrego AG. Integrating physi-
[41] Valde s CF, Chejne F, Marrugo GP, G?mez CA, Marin-Jaramillo A, Norena- cochemical information to follow the transformations of biomass upon tor-
Marin L. Energy evaluation of pelletized mixtures of IWTP sludge and coal- refaction and low-temperature carbonization. Fuel 2014;131:17e27. http://
fired boiler ashes by co-combustion: identification of synergistic effects. dx.doi.org/10.1016/j.fuel.2014.04.067.
Appl Therm Eng 2017;124:191e201. http://dx.doi.org/10.1016/ [68] Li M-F, Chen C-Z, Li X, Shen Y, Bian J, Sun R-C. Torrefaction of bamboo under
j.applthermaleng.2017.06.007. nitrogen atmosphere: influence of temperature and time on the structure and
[42] Granados DA, Vel asquez HI, Chejne F. Energetic and exergetic evaluation of properties of the solid product. Fuel 2015;161:193e6. http://dx.doi.org/
residual biomass in a torrefaction process. Energy 2014;74:181e9. http:// 10.1016/j.fuel.2015.08.052.
dx.doi.org/10.1016/j.energy.2014.05.046. [69] Ibrahim RHH, Darvell LI, Jones JM, Williams A. Physicochemical characteri-
[43] Colin B, Dirion JL, Arlabosse P, Salvador S. Experimental study of wood chips sation of torrefied biomass. J Anal Appl Pyrolysis 2013;103:21e30. http://
torrefaction in a pilot scale rotary kiln. Chem Eng Trans 2014;37:505e10. dx.doi.org/10.1016/j.jaap.2012.10.004.
http://dx.doi.org/10.3303/CET1437085. [70] Na II B, Ahn BJ, Lee JW. Changes in chemical and physical properties of yellow
[44] Basu P, Sadhukhan AK, Gupta P, Rao S, Dhungana A, Acharya B. An experi- poplar (Liriodendron tulipifera) during torrefaction. Wood Sci Technol
mental and theoretical investigation on torrefaction of a large wet wood 2014;49:257e72. http://dx.doi.org/10.1007/s00226-014-0697-1.
particle. Bioresour Technol 2014;159:215e22. http://dx.doi.org/10.1016/ [71] Toscano G, Pizzi A, Foppa Pedretti E, Rossini G, Ciceri G, Martignon G, et al.
j.biortech.2014.02.105. Torrefaction of tomato industry residues. Fuel 2015;143:89e97. http://
[45] Khan AA, de Jong W, Jansens PJ, Spliethoff H. Biomass combustion in fluidized dx.doi.org/10.1016/j.fuel.2014.11.039.
bed boilers: potential problems and remedies. Fuel Process Technol 2009;90: [72] Xu F, Yu J, Tesso T, Dowell F, Wang D. Qualitative and quantitative analysis of
21e50. http://dx.doi.org/10.1016/j.fuproc.2008.07.012. lignocellulosic biomass using infrared techniques: a mini-review. Appl Energy
[46] Yang H, Yan R, Chen H, Lee DH, Zheng C. Characteristics of hemicellulose, 2013;104:801e9. http://dx.doi.org/10.1016/j.apenergy.2012.12.019.
cellulose and lignin pyrolysis. Fuel 2007;86:1781e8. http://dx.doi.org/ [73] Pereira SC, Maehara L, Machado CMM, Farinas CS. Physical-chemical-
10.1016/j.fuel.2006.12.013. morphological characterization of the whole sugarcane lignocellulosic
[47] Wang G, Luo Y, Deng J, Kuang J, Zhang Y. Pretreatment of biomass by torre- biomass used for 2G ethanol production by spectroscopy and microscopy
faction. Chin Sci Bull 2011;56:1442e8. http://dx.doi.org/10.1007/s11434-010- techniques. Renew Energy 2016;87:607e17. http://dx.doi.org/10.1016/
4143-y. j.renene.2015.10.054.
[48] Deng J, Wang G, Kuang J, Zhang Y, Luo Y. Pretreatment of agricultural residues [74] Liu Q, Wang S, Zheng Y, Luo Z, Cen K. Mechanism study of wood lignin py-
for co-gasification via torrefaction. J Anal Appl Pyrolysis 2009;86:331e7. rolysis by using TGeFTIR analysis. J Anal Appl Pyrolysis 2008;82:170e7.
http://dx.doi.org/10.1016/j.jaap.2009.08.006. http://dx.doi.org/10.1016/j.jaap.2008.03.007.
[49] Shen DK, Gu S, Bridgwater AV. The thermal performance of the poly- [75] Ghaffar SH, Fan M. Structural analysis for lignin characteristics in biomass
saccharides extracted from hardwood: cellulose and hemicellulose. Carbo- straw. Biomass Bioenergy 2013;57:264e79. http://dx.doi.org/10.1016/
hydr Polym 2010;82:39e45. http://dx.doi.org/10.1016/j.carbpol.2010.04.018. j.biombioe.2013.07.015.
[50] Burnham AK. Global chemical kinetics of fossil fuels. 2017. http://dx.doi.org/ [76] Chang S, Zhao Z, Zheng A, He F, Huang Z, Li H. Characterization of products
10.1007/978-3-319-49634-4. from torrefaction of sprucewood and bagasse in an auger reactor. Energy and
[51] Seyler CA. Petrology and classification of coal: pts I & II. Proc South Wales Inst Fuels, vol. 26. American Chemical Society; 2012. p. 7009e17. http://
Eng 1938;53:254e327. dx.doi.org/10.1021/ef301048a.
[52] Sule I. Torrefaction behaviour of agricultural biomass. The University of [77] Jiang Z, Liu Z, Fei B, Cai Z, Yu Y, Liu X. The pyrolysis characteristics of moso
Guelph; 2012. bamboo. J Anal Appl Pyrolysis 2012;94:48e52. http://dx.doi.org/10.1016/
[53] Bergman PC, Boersma A, Zwart RW, Kiel JH. Torrefaction for biomass co-firing j.jaap.2011.10.010.
in existing coal-fired power stations. Energy Res Cent Neth ECN ECNC05013 [78] Pastorova I, Botto RE, Arisz PW, Boon JJ. Cellulose char structure: a combined
2005;71. doi:ECN-Ce;05e013. analytical Py-GC-MS, FTIR, and NMR study. Carbohydr Res 1994;262:27e47.
[54] Sandoval-Torres S, Jomaa W, Marc F, Puiggali J-R. Causes of color changes in http://dx.doi.org/10.1016/0008-6215(94)84003-2.
wood during drying. For Stud China 2010;12:167e75. http://dx.doi.org/ [79] McGrath TE, Chan WG, Hajaligol MR. Low temperature mechanism for the
10.1007/s11632-010-0404-8. formation of polycyclic aromatic hydrocarbons from the pyrolysis of cellulose.
[55] Sundqvist B. Colour changes and acid formation in wood during heating. Lulea J Anal Appl Pyrolysis 2003;66:51e70. http://dx.doi.org/10.1016/S0165-
University of Technology; 2004. 2370(02)00105-5.

Potrebbero piacerti anche