You are on page 1of 8

Bioresource Technology 274 (2019) 439–446

Contents lists available at ScienceDirect

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

Pyrolysis of blend (oil palm biomass and sawdust) biomass using TG-MS T

Arshad Adam Salema , Ryan Man Wai Ting, Yong Kuan Shang
Mechanical Engineering Discipline, School of Engineering, Monash University Malaysia, Jalan Lagoon Selatan, Bandar Sunway 47500, Selangor, Malaysia

G R A P H I C A L A B S T R A C T

A R T I C LE I N FO A B S T R A C T

Keywords: The aim of this study was to pyrolyze individual (oil palm shell, empty fruit bunch and sawdust) as well as blend
Biomass biomass in a thermogravimetric mass spectrometry (TG-MS) from room temperature to 800 °C at constant
Blend heating rate of 15 °C/min. The results showed that the onset TG temperature for blend biomass shifted slightly to
Pyrolysis lower values. Activation energy values were also found to decrease slightly after blending the biomass.
TG-MS
Interestingly, the MS spectra of selected gases (H2O CH4, H2O, C2H2, C2H4 or CO, CH2O, CH3OH, HCl, C3H6, CO2,
Kinetics
HCOOH, and C6H12) evolved from blend biomass showed decreased in the intensity as compared to their in-
Chemical
dividual biomass. Overall, the blend biomass showed synergy which provides ways to expand the possibility of
utilizing multiple feedstocks in one thermo-chemical system.

1. Introduction mitigate the problem of burning biomass openly or in uncontrolled


conditions which usually led to air pollution. In addition to this, oil
In Malaysia, about 5.4 million ha of the oil palm agricultural land palm biomass has the potential to replace the existing fossil fuel source
generate ∼80% of the biomass, making it one of the largest con- by developing a carbon neutral cycle.
tributors of lignocellulosic biomass in the world (Loh, 2017). For in- Pyrolysis is regarded as one of the most promising approaches to
stance, over 400 palm oil mills process about 82 million tonnes of oil produce renewable chemicals and products which can either totally or
palm fresh fruit bunches annually which generates about 33 million partially replace petroleum products (Pogaku et al., 2016). In past few
tonnes of biomass (Loh, 2017; Nizamuddin et al., 2015). Therefore, decades, large amount of research has been done on the pyrolysis of
Malaysian government, policy makers, and industries are looking to biomass. Most of these have used thermogravimetric analyzers (TGA) to
utilize biomass to produce value added products by implementing a study the fundamental characteristics of biomass pyrolysis, but this
National Biomass Policy 2020 (NBS, 2018). These policies are directly equipment in standalone can only provide the thermal behaviour and
related to the sustainability of the oil palm industry as well as to kinetics of the biomass pyrolysis. However, TGA when coupled with


Corresponding author.
E-mail address: arshad.salema@monash.edu (A.A. Salema).

https://doi.org/10.1016/j.biortech.2018.12.014
Received 17 October 2018; Received in revised form 5 December 2018; Accepted 6 December 2018
Available online 07 December 2018
0960-8524/ © 2018 Elsevier Ltd. All rights reserved.
A.A. Salema et al. Bioresource Technology 274 (2019) 439–446

mass spectrometry (MS) would enable better and detailed under- Table 1
standing of thermal, kinetics and chemical analysis of gases emitted Proximate and chemical analysis of biomass samples.
during pyrolysis (Fermoso and Mašek, 2018; Huang et al., 2018; Wang Proximate analysis (wt.%)
et al., 2017; Yang et al., 2004) and this can assist the large-scale pyr-
olysis plant operation (Rodilla et al., 2018; Wu et al., 2018). EFB OPS SD
Several researchers have pyrolyzed single biomass of different spe-
Moisture content 6.10 7.40 10.24
cies such as coffee ground residues (Fermoso and Mašek, 2018), castor Volatile matter 62.68 65.50 68.10
husk, castor stem, agave bagasse, coffee pulp, opuntia stem and pinus Ash 3.75 6.60 0.51
sawdust (Parascanu et al., 2017), olive tree pruning, olive kernel, vine Fixed carbon 27.47 20.50 21.15
shoots and grape husks (Sfakiotakis and Vamvuka, 2018) in a TG-MS Chemical content (wt.%)
was found. However, very few (Yang et al., 2004) or no study on the Cellulose 37.0a 20.5a 41.0b
pyrolysis of oil palm biomass using TG-MS is available. Therefore, the Hemicellulose 35.0a 22.5a 30.6b
Lignin 24.0a 52.0a 25.4b
pyrolysis mechanism of oil palm biomass is still not clear and there is
shortage of key information that is needed to optimize the key para- a
Shibata et al., 2008.
meters. b
Rafiqul and Sakinah, 2012.
Furthermore, those researchers who have carried out the pyrolysis
of blended biomass have mixed biomass either with coal, plastic or weighed using an AND GR-200 precision analytical balance with
sewage sludge. For instance, Otero et al., 2011 investigated thermal and 0.0001 g reading sensitivity and mixed appropriately with ratios of
chemical properties of manure and coal biomass blend via co-firing. 50:50 and 25:75. The ratio of 50–50 wt% and 25–50 wt% was selected
They found lower activation energy in the blend samples, which was to investigate the influence of mixing oil palm biomass with sawdust. In
also in agreement with the results of Idris et al., 2010, when oil palm typical power plant, sawdust and its pellet are used to produce heat and
biomass and sub-bituminous coal was blended. Wang et al., 2016, re- power as well as to produce some chemicals. Therefore, to investigate
ported that blend (seaweed and rice husk) biomass showed variation in the effect on the pyrolysis kinetics and gaseous product, oil palm bio-
the gaseous product composition as compared to pyrolysis of individual mass was mixed with sawdust in ratio of 50–50% and 25–75%. SD100,
biomass using different mixture ratios in a TG-MS. Jayaraman et al., EFB100, and OPS 100 refer to individual biomass as sawdust 100%,
2017 studied thermogravimetric and mass-spectrometric (TG-MS) of empty fruit bunch 100% and oil palm shell 100%, respectively.
coal-biomass blends under air atmosphere. Similarly, Rodilla et al., SD50 + OPS50 refer to blend biomass as 50 wt% sawdust mixed with
2018 analyzed the thermal, kinetics and chemical compounds of sub- 50 wt% oil palm shell, similarly SD25 + EFB75% refer to blend biomass
bituminous coal, energy crops and their blends in nitrogen, air and oxy- as 25 wt% sawdust mixed with 50 wt% oil palm shell. Mixing of all
fuel combustion atmospheres. Further, blending process would help in three biomass SD33 + EFB33 + OPS33 refer to 33 wt% sawdust mixed
investigating the blending ratio as well as selection of materials to mix with 33 wt% empty fruit bunch and 33 wt% oil palm shell. To ensure
with biomass (Rodilla et al., 2018). proper mixing, the biomass samples were subjected to an IKA Vortex 3
Very recently, Mallick et al., 2018 carried out pyrolysis in which vortex shaker at a speed of 2500 rpm for duration of 2 min.
two or more biomass such as saw dust, bamboo dust and rice husk were
blended to study the pyrolysis kinetics. However, in their study che-
2.2. Pyrolysis of blend biomass using TG-MS
mical analysis of the gaseous product using MS was not included. Be-
sides this study, none have focused on blending two or more types of
Pyrolysis of individual and blend biomass were carried out in a TA
biomass and pyrolyzing them in a TG-MS. It is expected that blending
Instruments Q500 thermogravimetric analyzer coupled to a TA
two or more biomasses might alter the pyrolysis kinetics as well as the
Instruments Discovery mass spectrometer (TG-MS). The gaseous pro-
chemical composition of the product. This technique also aims to pro-
duct generated from the pyrolysis were analyzed using a benchtop
vide an understanding of pyrolyzing multiple species blend biomass in a
quadrupole MS facility of 70 eV electron ionisation source and allowed
single thermo-chemical reactor.
mass spectra from 10 to 300 a.m.u. The evolved gases from TGA was
To our knowledge and based on the literature survey, this study
transferred to quadrapole MS via a heated capillary interface main-
presents for the first time the pyrolysis of two or more blend biomass
tained at 300 °C and was chemically resistance to avoid any con-
(oil palm shell, empty fruit bunch and sawdust) feedstock using a TG-
densation as well as contamination of the gaseous products.
MS. This paper also investigates the effect of blending ratio (between oil
Approximately 3 to 8 mg sample was loaded in the TGA and pryolyzed
palm biomass and sawdust) on thermal, kinetics and chemical char-
from room temperature to 800 °C at a heating rate of 15 °C/min. Pure
acteristics from room temperature to 800 °C at a constant heating rate
nitrogen (99.999%) was used as the carrier gas with a flow rate of
of 15 °C/min. This study might develop a foundation of investigating
200 ml/min. This study primarily identified the mass-to-charge ratios
the thermal and chemical characteristics of the blend biomass in a
(m/z) of 16, 18, 26, 28, 30, 31, 36, 42, 44, 46, and 84 a.m.u. which
scaled up design.
corresponds to CH4, H2O, C2H2, C2H4 or CO, CH2O, CH3OH, HCl, C3H6,
CO2, HCOOH, and C6H12, respectively were detected according to the
2. Methodology
database of National Institute of Standards and Technology (NIST).
2.1. Materials
2.3. Kinetics
The empty fruit bunch (EFB) and oil palm shell (OPS) biomass from
oil palm (Elaeis guineensis) plant were acquired from local oil palm mills The kinetics analysis of biomass pyrolysis was evaluated by Coats-
in Malaysia. Sawdust (SD) was acquired from a local wood furniture Redfern method (Sait et al., 2012). The general equation of reaction
processing facility which could be either from rubber tree or forest rate is given as:
timber. The biomass samples were dried at 105 °C for 24 h using an
dx
electric oven, crushed, grinded to particle size of 500–800 μm using a = kf (x )
(1)
dt
pulveriser and kept in air tight seal bags for further analysis. Table 1
presents the proximate analyses of biomass samples. The moisture where, k is the reaction rate constant and f (x ) refers to a selected
content was detected using oven-drying method while volatile matter model of reaction mechanism. x is the conversion degree of biomass
and fixed carbon was presented from the TGA data. The samples were which is expressed below in Eq. (2).

440
A.A. Salema et al. Bioresource Technology 274 (2019) 439–446

m o − mt of different sizes and shapes. Trombetta et al., 2010, also presented


x=
mo − m f (2) similar physical characteristics of SD. EFB seemed to have a rigid sur-
face covered by a layer of matrix material. This material is most
where, mt is corresponded to sample mass at time t , mo , and mf are the
probably the cuticle layer that prevents water loss in most plants.
initial and final mass of the sample respectively. Through the Arrhenius
Furthermore, numerous silica-bodies were found to be embedded to
equation, the reaction rate constant is obtained as:
circular craters which were almost uniformly spread over the surface.
−E The diameter of the round-shaped spikey silica-bodies was about 15 µm
k = exp ⎛ a ⎞
⎝ RT ⎠ (3) as reported by Law et al., 2007. Moreover, silicon and oxygen are the
where, A is the rate constant pre-exponential factor, Ea , R , and T are integral parts of the silica-bodies (Barros et al., 2003) as depicted in the
activation energy (kJ/mol), universal gas constant EDX mapping. Several uniformly dispersed micro pores were found on
(8.314 J.mol−1.K−1), and temperature (K) respectively. Furthermore, the surface of the OPS. The size of the micro pores were about 1.6 µm.
the reaction model in Eq. (1) can be shown as: The OPS are also irregular in shape and size and tend to agglomerate or
coat on the EFB and SD particles due to their rough fibrous surface
f (x ) = (1 − x )n (4) morphology. Thus, during pyrolysis the blend biomass particles would
where, n is the reaction order. Rearranging Eq. (1) and integrating it for undergo degradation simultaneously which might alter the kinetics and
a constant pyrolysis heating rate of β, where β = dT/dt: chemical properties of the process.

ln(1 − x ) ⎤ AR ⎛ 2RT ⎞ ⎤ E 3.2. Pyrolysis of individual and blend biomass


ln ⎡− = ln ⎡ 1− ⎜ − a ⎟

⎣ T2 ⎦ ⎢ βEa ⎝ Ea ⎠ ⎥ RT (5)
⎣ ⎦
2RT The decomposition of biomass in the present study was divided into
In Eq. (5), the value of ≪ 1, thus it is simplified to:
Ea four regions; Region I – drying (30–120 °C), Region II – low volatile
ln(1 − x ) ⎤ AR ⎤ E (120–220 °C), Region III – main volatile (220–400 °C), and Region IV –
ln ⎡− = ln ⎡ − a char (400–700 °C).
⎣ T2 ⎦ ⎢
⎣ βEa ⎥
⎦ RT (6)
The initial mass loss for all the samples occurred between room
ln(1 − x )
Therefore, the plot of ln ⎡− ⎤ versus 1 is linear. The slope and temperature and 120 °C (region I) as shown in Fig. 1 is mainly due to
⎣ T2 ⎦ T
Ea AR release of moisture or water evaporation. The mass loss in this region I
intercept are − and ln ⎡ βE ⎤ respectively. Hence, the pre-exponential
RT ⎣ a⎦ was about 2.0 wt% ± 0.2 for all samples (individual and their blend),
factor, A and activation energy, Ea can be calculated. This was achieved also depicted by a small peak (Peak 1) in the DTG curves as shown in
via plots of the following: Fig. 1.
1 − (1 − x)1 − n ⎤ 1 After the release of moisture, some light volatiles are expected to
ln ⎡− versus (for n ≠ 1) release in region II within the temperature range of 120–220 °C with

⎣ T 2 (1 − n) ⎥ ⎦ T (7)
negligible mass loss. Usually in this region the biomass chemical
or structure starts to depolymerize and soften. This region is not well ex-
ln(1 − x ) ⎤ 1 plained in most of the previous literature.
ln ⎡− versus (for n = 1) The major mass loss occurred in the region III, where a drastic drop
⎣ T2 ⎦ T (8)
in the mass took place due to the release of volatile matters between
The data of x and T was provided from TGA analysis to obtain the temperatures 220 and 400 °C. Table 2 shows the onset and peak tem-
Ea and A values. The criterion to acquire accurate and acceptable Ea perature along with the leftover residual for both individual and blend
and A values were based on the final value of n that should yield values biomass. An obvious difference between the individual and blend bio-
of Ea with highest coefficient of determination, R2 to the fitted regres- mass is observed in this region. For instance, the onset temperature
sion line. Initially, the reaction order was assumed to be first order for (Ton) for EFB (∼218 °C) biomass was earliest as compared to SD
the simplicity in calculation of kinetic parameters. Later, a numerical (∼222 °C) and OPS (∼231 °C). This might be due to the difference in
solver was used to determine the reaction order that gave the best linear the chemical structure and the lignocellulosic composition of the bio-
regression. Thermo-chemical degradation of biomass materials also mass (Wang et al., 2017) which is usually complex. The Ton was also
relies on strong connection between physical and chemical character- found to alter when the SD was blended with EFB and OPS biomass. As
istics (Sait et al., 2012). Hence, reaction performances can be analyzed seen in Fig. 1C, the Ton for SD and EFB blend biomass (50:50 wt% and
and anticipated from the kinetic analysis. 25:75 wt%) was almost between their parent or individual biomass.
Thus, blending biomass can either increase or decrease the onset tem-
2.4. FE-SEM with EDX perature showing a synergetic effect. Interestingly, the TG curve for all
(three) blended biomass (SD + EFB + OPS in Fig. 1I) was found to be
The physical characteristics of biomass was analyzed using FE-SEM the collective action of the individual biomass. However, the difference
(Hitachi SU8020 UHR Cold-Emission) and Oxford-Horiba Inca XMax50 in the TG curves was less pronounced in region III (pyrolysis) as com-
EDX. Samples were coated with 2–6 nm of gold coating using a sputter pared to region IV (char), which is in well agreement with previous
coater in order to minimize sample charging because of non-conducting study (Mallick et al., 2018).
materials. Images were taken at several magnifications. Beyond temperature 400 °C (region IV), a gradual decrease in the
mass loss was mainly due to slow pyrolysis of lignin component. The
3. Results and discussion final residues (char) at temperature 700 °C for individual biomass SD,
EFB, and OPS was 3.50, 3.72, and 13.00 wt%, respectively. However,
3.1. Physical and elemental analysis of biomass using FE-SEM and EDX the char for blend biomass (SD50 + EFB50, SD25 + EFB75,
SD50 + OPS50, SD25 + OPS75, and SD33 + EFB33 + OPS33 was
The FE-SEM micrographs for SD, EFB, OPS, and their blends 1.70, 6.40, 0.90, 0.22, and 7.82 wt%, respectively) decreased sig-
(SD33 + EFB33 + OPS33) is shown in Fig. S1. Typically, the surface nificant. This variation clearly indicates the interaction and synergetic
topography in the case of SD and EFB was rough, while OPS showed effect among the biomass during pyrolysis. It seems that the chemical
clear micro pores. EFB and SD are generally fibrous materials and OPS components or element composition in some of the biomass are acting
is a hard shell and non-fibrous material. The morphology of SD can be as catalyst. As observed in Rodilla et al., 2018, the char content de-
seen to have longitudinal type structure with surface defects and fibers creased significantly when biomass was blended with coal because

441
A.A. Salema et al. Bioresource Technology 274 (2019) 439–446

(caption on next page)


442
A.A. Salema et al. Bioresource Technology 274 (2019) 439–446

Fig. 1. TGA and DTG curves of (A) and (B) individual biomass, and (C) to (J) blend biomass at heating rate of 15 °C/min. (SD -sawdust, EFB – empty fruit bunch, OPS
– oil palm shell).

biomass has higher volatile matter than coal. According to this, significant difference in the thermal trend as compared to their parent
blending two or more biomasses in the present study might have in- biomass. The thermal behavior of SD + OPS blend mostly followed the
creased the volatile matter in a collaborative way to reduce the char individual SD biomass. Interestingly, EFB and OPS blends exhibited
content. significant changes in the thermal behaviour as depicted in Fig. 1G and
Among the individual biomass, OPS100 showed lowest degradation 1H. The derivative mass loss rate in the main pyrolysis region decreased
rate of about 8.0 wt.%/°C due to high composition of lignin and phy- drastically as well as two distinctive peaks was observed in this region.
sical nature of OPS. Further, EFB100 degraded much earlier at tem- This is due to strong interaction of lignocellulosic content in the blend
perature (325 °C) compared to OPS and SD, also the degradation rate biomass.
(9.8 wt.%/°C) was higher which might be due to its high cellulose The TG and DTG trend for all three blend biomass
content (Asadieraghi and Daud, 2015; Yang et al., 2004). The carbon (SD33 + EFB33 + OPS33) was also different from their parent in-
content is usually lower in EFB100 biomass as depicted in Table 3. dividual biomass (Fig. 1I and 1J). In this case, two distinct peaks were
Prominent and maximum degradation peaks at temperatures (325 °C revealed in the pyrolysis region similar to that of EFB + OPS blends,
for EFB100 and 375 °C for SD100 and OPS100) are generally due to and the derivative mass loss rate was significantly lower than individual
decomposition of cellulose materials. Moreover, no shoulder peak was biomass. Overall, the thermogravimetric trends of blend biomass were
observed in EFB100 biomass which is usually attributed to the de- mostly between the individual biomass which reveals possible syner-
composition of hemicellulose. This is consistent with previous literature getic interactions between the biomass as observed by Puig-Gamero
(Idris et al., 2010). However, the shoulder peak was significant in case et al., 2018.
of OPS100 and slightly in case of SD100. Table 2 shows the peak temperatures for individual and blend
Hemicellulose ([C5(H2O)4]n) consist of a random, amorphous biomass. The onset and end temperature was analyzed in TRIOS soft-
structure with weak strength and therefore its degradation is depicted ware using the intersection of an initial tangent line with a final tangent
much earlier (220–300 °C) in form of shoulder peak. In contrast, the line. The peak temperature was detected from the DTG curve corre-
cellulose ([C6(H2O)5]n) molecular structure consists of a long branch- sponding to the maximum degradation rate at the respective peak de-
less polymer chain of glucose units and usually forms a crystalline fined in Fig. 1B. A slight variation in these temperatures was observed
structure which makes it strong and resistant to heat. Therefore, in the blend biomass. The char residual percentage remained at 500 °C
thermal degradation of cellulose begins at higher temperatures i.e. did not showed any significant variation between individual and blend
beyond 320 °C and continues until 400 °C On the other hand, lignin biomass. However, such variation was significant at much higher
([C10H12O3]n) molecules are heavily cross-linked with superior strength temperature of 700 °C. This might be because of continuing pyrolysis of
and therefore, they degrade over a wider range of temperature staring biomass material at 500 °C with release of heavy chemical compounds
from 160 to 900 °C (Yang et al., 2004). According to Mishra and due to degradation of lignin. This can be also revealed from TGA curves
Mohanty, 2018, second and third peak in DTG curves are mainly due to (Fig. 1), whereby the weight loss is continued at 500 °C and beyond.
decomposition of hemicelluloses and cellulose components. The weight almost ceased to loss further for SD (700 °C) and EFB
There was an interesting and distinctive result of thermal behaviour (615 °C) materials, but it continued to drop for OPS even beyond
in the blend biomass. For instance, the TG and DTG curves of SD + EFB 800 °C.
blends (Fig. 1D) mostly followed the individual EFB biomass which
showed the dominance behavior of EFB over SD. Interestingly, the
derivative weight loss decreased in the main pyrolysis region for all 3.3. Kinetics analysis
blend biomass. Blending biomass also shifted the derivative peaks be-
cause of synergetic interaction between the biomasses. Also difference As expected, the lowest activation energy (Table 4) is attributed to
in peak height of DTG presents difference in reactivity of biomass the drying stage (removal of moisture) followed by the IV region (char
materials (Rodilla et al., 2018) with lower reactivity for blend biomass. formation), II region (low volatile matter), and lastly III region (main
Besides this, the shoulder peak was evident in SD50 + EFB50 (Fig. 1D) volatile reaction). The activation energy values for both individual and
blends but was absent in case of SD25 + EFB75 blends, which might be blend biomass samples in the main pyrolysis region (III) were in the
attributed to hemicelluloses content in SD. range of 101 to 110 kJ/mol. The relatively high activation energy in
Blend of SD and OPS (Fig. 1E and 1F) did not showed much this region implies higher temperature required to decompose the
biomass. The activation energy agreed with other studies (Asadieraghi

Table 2
Peak temperatures and residue leftover for individual and blend biomass.
Biomass Ton , °C Tend , °C Tpeak2 , °C Tpeak3 , °C Residue, wt.%

TG* TG DTG# DTG TG @500 °C TG @700 °C

SD 300 410 375 635 20.10 3.50


EFB 290 382 332 555 18.50 3.72
OPS 308 375 372 725 26.40 13.00
SD50 + EFB50 290 410 335 600 19.20 1.72
SD50 + OPS50 308 420 370 665 21.88 0.90
EFB50 + OPS50 284 410 332 600 22.00 2.36
SD25 + EFB75 290 390 334 625 23.60 6.40
SD25 + OPS75 305 420 370 675 22.30 0.22
EFB25 + OPS75 280 380 335 370 26.00 10.36
SD33 + EFB33 + OPS33 288 390 336 368 25.00 7.82

* TG – Thermogravimetric.
#
DTG – Derivative thermogravimetric.

443
A.A. Salema et al. Bioresource Technology 274 (2019) 439–446

Table 3
Elemental analysis of individual and blend biomass using FE-SEM-EDX.
Element, wt.% SD EFB OPS SD50 + OPS50 SD33 + EFB33 + OPS33

Carbon 70.66 65.71 69.88 55.31 53.13


Oxygen 15.29 15.95 14.04 44.69 41.17
Calcium 0.71 – – – –
Silicon – 8.71 – – –
Potassium – 2.73 – – –
Iron – 0.17 – – –
Chlorine – 0.78 – – –

and Daud, 2014; Asadieraghi and Daud, 2015; Yang et al., 2004). CH3 OH → thermal decomposition ( > 400 °C) → CH4
Generally, activation energy in the char region is lower, which mainly
consists of lignin pyrolysis (Yang et al., 2004). The coefficient of de- Water (H2O) corresponding to m/z ratio of 18 is produced at two
termination (R2 ) was in the range of 0.956–0.997 which signified best stages. Firstly, the physically absorbed waster in form of moisture is
fitting of the data. released at lower temperature of ∼75 °C till 150 °C. Secondly, sig-
Overall, the activation energy is lower in blend biomass than in- nificant amount of reaction water is produced during pyrolysis mainly
dividual. This result is consistent with previous studies (Jayaraman in temperature range of 250–400 °C, due to decomposition of cellulose
et al., 2017; Mallick et al., 2018; Rodilla et al., 2018). It is difficult to (Wang et al., 2017). However, a continuous increasing trend of water
investigate the exact cause of decrease and variation of activation en- production in both the individual and blend biomass was observed
ergy, but as a general rule of thumb such variation is a result of var- during pyrolysis. This is because the MS instrument is constantly under
iation in reaction mechanism (Rodilla et al., 2018) which is practically vacuum, if the TGA output flow is not the same as the MS entry vacuum
complex to predict. Besides this, the change in activation energy after flow, a small part of the ambient atmosphere might catch due to high
blending is also consequences of accelerated heating due to high vo- ambient humidity and therefore a continual trend of water might have
latile and catalytic effect by the biomass composition (Mallick et al., been detected by MS. The upward trend of the mass spectra of H2O after
2018). However, with the increase in OPS fraction in the blend, the temperature over 220 °C might also form during the thermal decom-
activation energy values were found to increase as well. This is because position of hemicellulose, cellulose, and lignin (Gomez et al., 2007).
of high activation energy needed to pyrolyze the OPS biomass. Similar However, high production of water is undesirable because it deterio-
effect on the activation energy was observed (Jayaraman et al., 2017) rates the quality of bio-oil (Wang et al., 2017). Large amounts of H2O
in coal-biomass blending. and CO were detected because of high content of hydroxyl groups and
oxygen atoms that are present in the biomass samples (Malika et al.,
3.4. Chemical analysis 2016).
The MS spectra for CO, ethylene (C2H4), and formaldehyde (CH2O)
As indicated in Fig. 2, most of the gases evolved showed distinctive also did not follow the DTG curve. The oxygenated group such as car-
changes in trend at a temperature range of 200–700 °C. The detected boxyl and carbonyl groups, may decompose into chemicals, such as
gases evolved in stages of mass spectra were barely consistent with the CO2, CO and formaldehyde (Wang et al., 2017). Gases such as acetylene
stages of TG and DTG. The most obvious gases detected in all the cases (C2H2) methanol (CH3OH), HCl, propene (C3H6), CO2, formic acid
were carbon dioxide (CO2), methanol, propene, and formic acid. Most (CH2O2), and cyclohexane (C6H12) coincident fairly with the TG and
of the gases are believed to release in a narrow temperature range of DTG profiles. These gases showed greatest MS signal during the main
300–450 °C. pyrolysis region III (Fig. 2) in the temperature range of 250–750 °C.
Evolution of gases occur mainly via two reactions namely, gas-phase Among these gases, methanol, propene, CO2, and formic acid depicted
homogeneous and gas-solid heterogeneous reactions. A slight increase clear peaks which stretched between 250 °C and 700 °C. Patwardhan
in trend of methane (CH4) production corresponding to the m/z ratio of et al., 2011, reported that decomposition of hemicellulose produces
16 was detected beyond ∼450 °C temperature in case of individual CO2 at 250 °C and it increases with temperature, formic acid was de-
biomass, but the increase was significant in case of blend biomass and tected from 350 °C to 600 °C. Typically, CO2 is the product from the
started to produce at much lower temperature of ∼300 °C. It has been decomposition of hemicellulose, cellulose, and lignin pyrolysis. Fig. 2
reported (Özveren and Özdoğan, 2013; Wang et al., 2017) that de- illustrates that there were 2 peaks in the mass spectra of SD100 at m/z
composition of methoxyl group can lead to the formation of smaller ratio of 44 (CO2) at 490 °C and 680 °C, respectively. This was consistent
compounds such as CH4 above 400 °C. For e.g. as follow with the DTG peaks of SD100 which indicated major reactions taking

Table 4
Kinetic parameters of individual and blend biomass samples.

1- Drying region.
2 - Light volatile region.
3 - Main volatile region.
4 - Char region.

444
A.A. Salema et al. Bioresource Technology 274 (2019) 439–446

pyrolysis reactions and formation of compounds at different stages of


pyrolysis. Earlier studies have also found similar common non-con-
densable gases such as CO, CO2 and CH4, and condensable volatiles,
such as H2O, methanol, acids and phenols during biomass pyrolysis
using TG-FTIR and TG-MS analyzer (Wang et al., 2017). The difference
in the MS signals in form of intensity of ion current (A) between in-
dividual and blend biomass presents a clear synergetic effect. An initial
drop in the MS signals was observed in the blend biomass from room
temperature until ∼200 °C for methane, acetylene, CO and C2H4, for-
maldehyde, methanol, and CO2 gases which was not significant in in-
dividual biomass. Moreover, the temperature at which the gases were
released also differed in blend biomass. It is also expected that the
quantity of gases might also get affected due to blending process. The
MS signals for HCl, acetylene, and cyclohexane gas in blend biomass
presented large fluctuations similar to individual biomass. On the other
hand, due to the chances of atom or molecule interferences that have
identical atomic or molecular weights, CH4 and C2H4 were represented
by m/z ratios of 15 and 27 instead of 16 and 28 respectively. Further-
more, the mass spectra of SD25 + OPS75 was found to be different from
most of the other biomass trends. The trends showed increasing pro-
ductivity of the evolved gases compared to its individual t biomass
(SD100). This was likely due to synergistic effects from the reaction of
evolved gases of two different biomass samples.

4. Conclusions

Pyrolysis of blend biomass can alter both the thermal as well as the
chemical product due to synergetic interactions. For instance, the ac-
tivation energy for blend biomass was lower than the individual bio-
mass which can decrease the initial or onset temperature during pyr-
olysis. Main evolved gases such as CO2, CH4, Methanol, C3H6, and
CH2O2 varied in intensity with increase in the temperature. The sy-
nergetic effect signifies that different types of biomass can be blended
and pyrolyzed in one reactor system which may avoid capital and op-
eration cost.

Acknowledgements

The authors would like to thank NanoQAM Research Center,


Department of Chemistry, University of Quebec at Montreal, Canada for
carrying out TG-MS analysis of the biomass samples. We acknowledge
the financial support from Discipline of Mechanical Engineering,
Monash University Malaysia.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://


doi.org/10.1016/j.biortech.2018.12.014.

References

Fig. 2. Chemical compounds from individual and blend biomass during pyr- Asadieraghi, M., Daud, W.M.A.W., 2014. Characterization of lignocellulosic biomass
thermal degradation and physiochemical structure: effects of demineralization by
olysis. (SD – sawdust, EFB – empty fruit bunch, OPS – oil palm shell).
diverse acid solutions. Energy Convers. Manage. 82, 71–82.
Asadieraghi, M., Daud, W.M.A.W., 2015. In-depth investigation on thermochemical
characteristics of palm oil biomasses as potential biofuel sources. J. Anal. Appl. Pyrol.
place within this temperature range.
115, 379–391.
The MS signal for methanol, propene, and CO2, gases were max- Barros, C.F., da Cunha, M., Miguens, F.C., Lins, U., 2003. Silica biomineralization in the
imum in case of individual biomass at temperature ∼450 °C, 500 °C, palm tree Syagrus coronata (Mart.) Becc.: structure, morphology and composition of
and 500 °C, respectively. The peak signal for formic acid depended on silicon biocomposites. Microsc. Microanal. 9 (S02), 1522–1523.
Fermoso, J., Mašek, O., 2018. Thermochemical decomposition of coffee ground residues
the types of biomass. For instance, EFB biomass showed one formic acid by TG-MS: a kinetic study. J. Anal. Appl. Pyrol. 130, 358–367.
peak at 600 °C, while OPS showed two clear peaks at 500 and 700 °C. Gomez, C.J., Meszaros, E., Jakab, E., Velo, E., Puigjaner, L., 2007. Thermogravimetry/
There appeared two peaks in case of blend SD50 + OPS50 and mass spectrometry study of woody residues and an herbaceous biomass crop using
PCA techniques. J. Anal. Appl. Pyrol. 80 (2), 416–426.
SD25 + OPS75 biomass in compare to SD50 + EFB50 and Huang, L., Xie, C., Liu, J., Zhang, X., Chang, K., Kuo, J., Sun, J., Xie, W., Zheng, L., Sun, S.,
SD25 + EFB75 blend biomass which showed only one peak. The MS Buyukada, M., Evrendilek, F., Buyukada, M., 2018. Influence of catalysts on co-
signals for acetylene, HCl, and cyclohexane were observed fluctuating combustion of sewage sludge and water hyacinth blends as determined by TG-MS
analysis. Bioresour. Technol. 247, 217–225.
between temperature 100 and 750 °C. This could be due to complex Idris, S.S., Rahman, N.A., Ismail, K., Alias, A.B., Rashid, Z.A., Aris, M.J., 2010.

445
A.A. Salema et al. Bioresource Technology 274 (2019) 439–446

Investigation on thermochemical behaviour of low rank Malaysian coal, oil palm Puig-Gamero, M., Lara-Díaz, J., Valverde, J.L., Sánchez, P., Sanchez-Silva, L., 2018.
biomass and their blends during pyrolysis via thermogravimetric analysis (TGA). Synergestic effect in the steam co-gasification of olive pomace, coal and petcoke:
Bioresour. Technol. 101 (12), 4584–4592. Thermogravimetric-mass spectrometric analysis. Energy Convers. Manage. 159,
Jayaraman, K., Kok, M.V., Gokalp, I., 2017. Thermogravimetric and mass spectrometric 140–150.
(TG-MS) analysis and kinetics of coal-biomass blends. Renew. Energy 101, 293–300. Rafiqul, I.S.M., Sakinah, A.M., 2012. Design of process parameters for the production of
Law, K.N., Daud, W.R.W., Ghazali, A., 2007. Morphological and chemical nature of fiber xylose from wood sawdust. Chem. Eng. Res. Des. 90 (9), 1307–1312.
strands of oil palm empty-fruit-bunch (OPEFB). BioResources 2 (3), 351–362. Rodilla, I., Contreras, M.L., Bahillo, A., 2018. Thermogravimetric and mass spectrometric
Loh, S.K., 2017. The potential of the Malaysian oil palm biomass as a renewable energy (TG-MS) analysis of sub-bituminous coal-energy crops blends in N 2, air and CO 2/O
source. Energy Convers. Manage. 141, 285–298. 2 atmospheres. Fuel 215, 506–514.
Malika, A., Jacques, N., Fatima, B., Mohammed, A., 2016. Pyrolysis investigation of food Sait, H.H., Hussain, A., Salema, A.A., Ani, F.N., 2012. Pyrolysis and combustion kinetics
wastes by TG-MS-DSC technique. Biomass Convers. Biorefin. 6 (2), 161–172. of date palm biomass using thermogravimetric analysis. Bioresour. Technol. 118,
Mallick, D., Poddar, M.K., Mahanta, P., Moholkar, V.S., 2018. Discernment of synergism 382–389.
in pyrolysis of biomass blends using thermogravimetric analysis. Bioresour. Technol. Sfakiotakis, S., Vamvuka, D., 2018. Thermal decomposition behavior, characterization
261, 294–305. and evaluation of pyrolysis products of agricultural wastes. J. Energy Inst. 91 (6),
Mishra, R.K., Mohanty, K., 2018. Pyrolysis kinetics and thermal behavior of waste saw- 951–961.
dust biomass using thermogravimetric analysis. Bioresour. Technol. 251, 63–74. Shibata, M., Varman, M., Tono, Y., Miyafuji, H., Saka, S., 2008. Characterization in
NBS, 2018. National Biomass Strategy 2020. accessed on November 28, 2018. http:// chemical composition of the oil palm (Elaeis guineensis). J. Japan Inst. Energy 87 (5),
www.nbs2020.gov.my/. 383–388.
Nizamuddin, S., Jayakumar, N.S., Sahu, J.N., Ganesan, P., Bhutto, A.W., Mubarak, N.M., Trombetta, E., Flores-Sahagun, T., Satyanarayana, K.G., 2010. Evaluation of poly-
2015. Hydrothermal carbonization of oil palm shell. Korean J. Chem. Eng. 32 (9), propylene/saw dust composites prepared with maleated polypropylene (MAPP)
1789–1797. produced by reactive extrusion. Matéria (Rio de Janeiro) 15 (2), 309–318.
Otero, M., Sánchez, M.E., Gómez, X., 2011. Co-firing of coal and manure biomass: a Wang, S., Dai, G., Yang, H., Luo, Z., 2017. Lignocellulosic biomass pyrolysis mechanism: a
TG–MS approach. Bioresour. Technol. 102 (17), 8304–8309. state-of-the-art review. Prog. Energy Combust. Sci. 62, 33–86.
Özveren, U., Özdoğan, Z.S., 2013. Investigation of the slow pyrolysis kinetics of olive oil Wang, S., Hu, Y., Wang, Q., Xu, S., Lin, X., Ji, H., Zhang, Z., 2016. TG–FTIR–MS analysis
pomace using thermo-gravimetric analysis coupled with mass spectrometry. Biomass of the pyrolysis of blended seaweed and rice husk. J. Therm. Anal. Calorim. 126 (3),
Bioenergy 58, 168–179. 1689–1702.
Parascanu, M.M., Sandoval-Salas, F., Soreanu, G., Valverde, J.L., Sanchez-Silva, L., 2017. Wu, X., Ba, Y., Wang, X., Niu, M., Fang, K., 2018. Evolved gas analysis and slow pyrolysis
Valorization of Mexican biomasses through pyrolysis, combustion and gasification mechanism of bamboo by thermogravimetric analysis, Fourier transform infrared
processes. Renew. Sust. Energy Rev. 71, 509–522. spectroscopy and gas chromatography-mass spectrometry. Bioresour. Technol. 266,
Patwardhan, P.R., Brown, R.C., Shanks, B.H., 2011. Product distribution from the fast 407–412.
pyrolysis of hemicellulose. ChemSusChem 4 (5), 636–643. Yang, H., Yan, R., Chin, T., Liang, D.T., Chen, H., Zheng, C., 2004. Thermogravimetric
Pogaku, R., Hardinge, B.S., Vuthaluru, H., Amir, H.A., 2016. Production of bio-oil from oil analysis− Fourier transform infrared analysis of palm oil waste pyrolysis. Energy
palm empty fruit bunch by catalytic fast pyrolysis: a review. Biofuels 7 (6), 647–660. Fuels 18 (6), 1814–1821.

446

You might also like