Sei sulla pagina 1di 16

Applied Energy 208 (2017) 142–157

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Systematic investigation on combustion characteristics and emission- T


reduction mechanism of potentially toxic elements in biomass- and biochar-
coal co-combustion systems

Balal Yousafa,b, Guijian Liua,b, , Qumber Abbasa, Ruwei Wanga, Muhammad Ubaid Alia,
Habib Ullaha, Ruijia Liua, Chuncai Zhoua
a
CAS-Key Laboratory of Crust-Mantle Materials and the Environments, School of Earth and Space Sciences, University of Science and Technology of China, Hefei 230026,
PR China
b
State Key Laboratory of Loess and Quaternary Geology, Institute of Earth Environment, The Chinese Academy of Sciences, Xi’an, Shaanxi 710075, China

H I G H L I G H T S

• The study investigates thermal-characteristics and migration of PTEs during biochar-coal combustion.
• Biochar-coal blends improved thermal behavior compare to coal/biomass-coal fuels.
• Soot yield and un-burned C in fly ash reduced significantly in biochar blended fuels.
• Volatilization of PTEs during co-firing of biochar-coal blends reduced considerably.
• Biochar-coal blend can be used in power plants for emission-reduction of PTEs.

A R T I C L E I N F O A B S T R A C T

Keywords: Thermochemically converted biochar is considered as one of the promising alternative solid-fuel due to its high
Biochar-coal blend carbon contents of up to 80%, and has great potential to produce environmentally-friendly green-energy by
Combustion characteristics improved fuel properties and emission-reduction of potentially toxic elements (PTEs). In this study, the biochar
Emission-reduction fuels, produced from peanut shell (PS) and wheat straw (WS) at 300, 500 and 700 °C, alone and blended with
Potentially toxic elements
coal at mass ratio of 20% and 50% were systematically investigated for combustion characteristics and their
Co-combustion system
potential to reduce the emission of PTEs including As, B, Ba, Be, Bi, Cd, Co, Cr, Cu, Ga, Ni, Pb, Sb, Sn, V and Zn in
relation to partitioning, retention and volatilization in the co-combustion systems, using a variety of experi-
mental techniques. Results indicated that the biochar-coal blended fuels in equal proportion showed steady state
combustion over broad temperature range resulting increased the combustion efficiency and improved the
thermal characteristics in comparison to coal and/or biomass-coal fuels. In addition, soot yield, CO emission and
un-burned carbon in fly ash reduced significantly in biochar-blended fuels. However, CO2 emission from bio-
char-coal co-combustion was comparable to coal and/or biomass-coal fuels. Moreover, the present study illu-
strated that the volatilization potential of PTEs during combustion of biochar and their blends with coal de-
creased considerably up to 21% compared to that of coal, and enrichment of these contaminants occurred in the
bottom and fly ashes ranged from 15.38–65% and 24.54–74.29%, respectively. Slagging and fouling problems
were still found with biochar-coal co-combustion due to the higher inorganic fraction of biochar, which were
overcome with the hydrothermal washing of fuels. Thus, it can be concluded that biochar-coal co-combustion is
a suitable option for its use in existing coal-fired energy generation system to achieve the sustainable clean-green
energy and reduction of gaseous PTEs emission.


Corresponding author at: CAS-Key Laboratory of Crust-Mantle Materials and the Environments, School of Earth and Space Sciences, University of Science and Technology of China,
Hefei 230026, PR China.
E-mail addresses: balal@ustc.edu.cn (B. Yousaf), lgj@ustc.edu.cn (G. Liu), qumber@mail.ustc.edu.cn (Q. Abbas), wrw@ustc.edu.com (R. Wang),
ubaid@mail.ustc.edu.cn (M. Ubaid Ali), habib901@mail.ustc.edu.cn (H. Ullah), nrj661@mail.ustc.edu.cn (R. Liu), zhoucc@ustc.edu.cn (C. Zhou).

http://dx.doi.org/10.1016/j.apenergy.2017.10.059
Received 6 May 2017; Received in revised form 13 October 2017; Accepted 14 October 2017
Available online 21 October 2017
0306-2619/ © 2017 Elsevier Ltd. All rights reserved.
B. Yousaf et al. Applied Energy 208 (2017) 142–157

1. Introduction oil solid fuel (biochar), syngas and variety of other valuable chemicals
depending upon process operating conditions such as highest heating
The use of renewable energy resources as an alternative to con- temperature (HHT), retention time, heating and sweep gas flow rates,
ventional non-renewable fossil fuel resources has currently received and pyrolysis environment [14,22–24]. In a comparison of raw biomass
attention worldwide [1]. This transition is primarily due to depletion of feedstock, the pyrolytic product biochar has improved fuel properties
fossil fuels and environmental concerns such as greenhouse gases such as higher energy density, higher calorific value, high carbon and
(GHGs) emission and release of potentially toxic elements through flue low oxygen content, and less emission of GHGs [8,25–26]. Moreover,
gas during coal combustion. The potentially toxic elements are any pyrolysis also immobilized the PTEs in biochar due to higher fixed
metallic elements (As, Ba, Bi, Cd, Cr, Cu, Ga, Ni, Pb, Sb, Sn, Zn, etc.) carbon and alkaline nature, which may reduce the PTEs emission
having the potential to cause severe health hazards if present in ex- during combustion process [27–28]. Thus, co-combustion of biochar
cessive concentration. Moreover, their emission in the environment is with coal in existing power plants likely results multiple, technological,
an issue of key concern nowadays due to their persistence in nature and economic and environmental benefits.
chemical stability in both aquatic and terrestrial ecosystems [2]. Fur- Knowledge about the combustion and kinetic characteristics of in-
thermore, chronic exposure of these PTEs to living organisms, present dividual or blended fuels are crucial for the designing of co-fired energy
in the environment, has accumulation potential in various body parts generation system [10]. Thermo-gravimetric analysis (TGA) is an im-
and can cause different abnormalities in humans [3–4]. portant and most commonly employed technique to investigate the
Currently, renewable energy meets ∼16% primary energy demands thermal behavior and kinetics of solid fuels such as biomass, coal and
on a global scale [5]. Renewable resources, particularly biomass ma- their blends in different proportion [9]. It enables to measure the rate of
terials, are the fastest growing energy source on earth system according change of sample mass with time as the function of temperature.
to current International Energy Outlook report [6]. For example, the Thermal decomposition and combustion stages temperatures can also
generation of renewable electricity in the USA from biomass is pro- be determined using TGA. Furthermore, kinetic parameters such as
jected to be triple by 2040. As plant materials have fixed carbon in activation energy of fuels individual combustion stages could be cal-
lignocellulose form during their growth by photosynthesis process and culated [10,29].
thus these are carbon negative due to no net contribution in greenhouse The availability of low-cost, good quality biomass material is a
effect with combustion. Thus, in comparison to stand-alone biomass prerequisite for commercial-scale production and use of biochar in the
combustion, biomass-coal blends co-utilization in existing coal-fired co-firing system [23]. Peanut shell (PS) and Wheat straw (WS) biomass
power plants is appealing these days because no additional cost or materials are abundantly available agricultural crop residues used in
substantial investment is required [7]. Biomass-coal combustion also this study. Efforts to investigate the combustion characteristics and
causes NOx and SO2 gaseous pollutants emission reduction due to low thermal behaviors of various biomasses under co-combustion system
sulfur (S) and higher volatile contents of biomass [8]. Additionally, co- have focused extensively in previous studies. However, limited studies
combustion is an effective approach to disposing of the crop residuse, address biochar importance as a solid fuel to improve both thermal
otherwise, on farm burning generates particulate matter (5–12%) that behaviors and emission-reduction of PTEs during co-combustion
deteriorate air quality and caused severe issues recently [9]. Extensive system. The overall objectives of this study were (1) to assess the in-
research on co-firing of different lingo-cellulose biomass materials such fluence of pyrolysis temperature on fuel properties of the peanut shell
as brewer's spent grains, wheat straw, rice husk and pine sawdust, with and wheat straw derived biochars. (2) to investigate the thermal be-
coal has been carried out recently [10–13]. havior and reaction kinetics of raw biomass, biochars, coal and their
However, there are several technical challenges associated with blends during combustion by using the thermogravimetric analyzer. (3)
biomass intrinsic properties that limit its application in existing co-fired The effect of pyrolysis on the partitioning behavior of PTEs in different
energy production system. These issues are higher contents of alkali- phases and their emission characteristics (volatilization potential)
alkaline earth metals and halogens in biomass materials that cause during coal/pyrolytic products co-utilization in an energy generation
fouling, slagging or corrosion problems in boiler during operation system.
[14–15]. Other drawbacks are higher oxygen contents resulting low
energy density caused flame instability during co-combustion, poor 2. Material and methods
grindability due to fibrous structure thus more energy is required for
grinding process as well as higher moisture and volatile matter contents 2.1. Collection and preparation of materials
of feedstock biomass [16]. All these concerns vary with biomass type
and their physico-chemical properties; limit the potential economic The Bituminous coal (C) having low ash content (∼13%), selected
utility of biomass materials for energy generation due to poor thermal for use in this study, obtained from Huainan Coalfield (Anhui, China)
efficacy and more air pollutants emission [17–18]. Moreover, the co- that mostly utilized for the generation of electricity in coal-fired
utilization of biomass with coal can redistribute PTEs in different Pingwei power station. Abundantly available agricultural waste bio-
phases such as in bottom ash, fly ash and flue gas. The processes in- mass materials, peanut shells (PS) and wheat straws (WS) were chosen,
volved in PTEs speciation are crucial as these influence the emission obtained from the Anhui province of China for this study. The coal and
characteristics of PTEs. The partitioning and migration behavior of biomass raw material samples were air dried, milled with a laboratory
PTEs in different combustion products vary with vaporization/con- scale grinder (Model 4 Wiley® Thomas Scientific, USA) and sieved to
densation potential of particular element volatile species, initial con- attain a homogenized particle size fraction in the range of 50–100
tent, occurrence modes such as physisorption or chemisorption, fuel mesh.
type, and air pollution emission control devices, as well as design and
operational conditions of combustion system [2]. Emission behavior of 2.2. Biomass pyrolysis and experimental process
PTEs during coal/biomass and their blends combustion have been fo-
cused recent years [19–21]. Thus, to overcome all these downsides and The schematic diagram of the fixed bed split furnace system (model
meet the challenges of existing coal fired power generation system it is BTF-1200C, Anhui BEQ equipment, technology Co., Ltd, China) used in
necessary to upgrade the fuel properties of biomass prior to use in the this study is illustrated in Fig. 1. Slow-pyrolysis was conducted at 300,
co-utilization system. 500 and 700 °C by taking approximately 30 g of air-dried and ground
In this regard, pyrolysis is an attractive thermochemical conversion biomass material in ultra-high purity quartz boat and placed inside the
technology to improve the biomass fuel properties and restricts the center of the reactor tube. The residual air was purged by using a
release of PTEs by stabilizing these in biochar. Pyrolytic yields, i.e., bio- constant supply of argon from one side of the tube with 50 ml min−1

143
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Fig. 1. Schematic diagram illustrating laboratory scale fixed bed combustion system and sampling procedure.

flow rate adjusted by a mass flow controller (MFC) to provide the inert 2.4. Thermo-gravimetric analysis
atmosphere. Ceramic tube blocks were put on both sides of sample
quartz boat to block the heat radiation loss from tube chamber. An The thermochemical behavior of feed coal, biomass materials, bio-
automatic PID controller (proportional-integral-derivative) integrated chars and their blends was studied under air atmosphere using a si-
into a furnace heated the biomass sample at a heating rate of 10 °C/min. multaneous thermal analyzer SDT Q600 (TA Instruments, USA)
When the highest heating temperature (HHT) was attained, the sample equipped with data acquisition system. This unit has temperature
was kept under constant conditions for 60 min. Prior to removing the measurement accuracy of 0.001 °C, the micro-balance sensitivity of
biochar sample from the furnace chamber, it was cooled to room ∼0.1 mg with wide heating rates ranging from 1 to 100 °C min−1. The
temperature. The detailed information about pyrolytic yield (biochar, calibration of instrument furnace was performed by using certified
bio-oil, gas, and liquid) at various temperatures and their chemical standard reference materials (indium/nickel) to ensure the accuracy of
compositions are given in Table S1–S4. the analytical method. To create good contact between sample and
A total of 25 samples including one coal (C), two biomasses crucible about 15 ± 1 mg powdered sample (< 50 µm or 200 mesh-
(PS & WS) and their biochars at various temperatures (300, 500 and size) was packed in an aluminum pan and thermally treated at ambient
700 °C) and sixteen (16) samples of coal-biomass and coal-biochar room temperature to 1000 °C at a heating rate of 20 °C min−1 and kept
blends were prepared by physically mixing the materials using porce- until constant weight loss. A continuous oxidized condition (100 ml air
lain mortar and pestle in different ratios. The detailed description about min−1: ∼21.05 ml oxygen (O2) min−1 and ∼78.95 ml nitrogen (N2)
the experimental tests is presented in Table 1. min−1 as a carrier gas) maintained during the thermal analysis of fuels.
As the function of temperature increase, the weight loss (thermo-
2.3. Compositional, proximate and ultimate analysis gravimetric) and weight loss rate (%/°C) (differential thermo-gravi-
metric) signals were recorded continuously during analysis. Three re-
The compositional analysis (cellulose, hemicellulose, and lignin) of plicates of each TGA experiment were performed to ensure the re-
biomass feedstock was performed by using the methodology of producibility.
American Society for Testing and Materials (ASTM). Characterization of
coal, raw feedstock and produced biochar regarding proximate analysis
2.5. Kinetic modeling and validation
was also conducted by following the ASTM International standard
methods.
The co-combustion of coal-biomass and coal-biochar blends is a
The percentage elemental (carbon, hydrogen, nitrogen, and sulfur)
multiple-step process as these are complex fuels having a broad range of
composition of the samples were determined by using vario MACRO
structural-chemical composition. Thus various reactions co-exist at
cube (Elementar, Germany) elemental analyzer. The oxygen content
different rates during this process that complicates the formulation of
(%) was calculated by subtracting the C, H, N, S, and ash contents from
reaction mechanism and kinetics [18,30]. An insight information about
100.
combustion system and optimum conditions can be achieved from the
The experimental calorific values (CV) as Higher Heating Values
chemical kinetic procedure. In the current study, for the analysis of
(HHV) of coal-biomass and coal-biochar blends were measured by uti-
non-isothermal TG (thermo-gravimetric) data basic method was applied
lizing oxygen bomb calorimeter model C2000 (IKA®, Germany). In
to calculate the kinetics of the dynamic thermo-chemical reaction.
addition, the calorific values (CV) of the samples were theoretically
As combustion is a complex multi-phase reaction, therefore, we
calculated by using the modified Dulong’s formula expressed in the
used Coats-Redfern simulation model for activation energy calculation
following Eq. (1).
[10,17]. The equation stated below was used to express the kinetic
33.5 × wt% C 142.3 × wt% H 15.4 × wt% O reaction rate:
HHV (MJ kg−1) = + −
100 100 100
dx / dt = kf(x) (2)
(1)

144
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Table 1
Description of experimental tests conducted.

Test no Test code Coal Biomass Biochar 300 Biochar 500 Biochar 700

T1 C 100 – – – –
T2 C80PS20 80 20 – – –
T3 C50PS50 50 50 – – –
T4 PS – 100 – – –
T5 C80PSBC300-20 80 – 20 – –
T6 C50PSBC300-50 50 – 50 – –
T7 PSBC300 – – 100 – –
T8 C80PSBC500-20 80 – – 20 –
T9 C50PSBC500-50 50 – – 50 –
T10 PSBC500 – – – 100 –
T11 C80PSBC700-20 80 – – – 20
T12 C50PSBC700-50 50 – – – 50
T13 PSBC700 – – – – 100
T14 C80WS20 80 20 – – –
T15 C50WS50 50 50 – – –
T16 WS – 100 – – –
T17 C80WSBC300-20 80 – 20 – –
T18 C50WSBC300-50 50 – 50 – –
T19 WSBC300 – – 100 – –
T20 C80WSBC500-20 80 – – 20 –
T21 C50WSBC500-50 50 – – 50 –
T22 WSBC500 – – – 100 –
T23 C80WSBC700-20 80 – – – 20
T24 C50WSBC700-50 50 – – – 50
T25 WSBC700 – – – – 100

Ea different heating rates. To apply the comprehensive method, the Coats-


While, k= Aexp ⎛− ⎞
⎝ RT ⎠ (3) Redfern (CR) model, well known for activation energy calculation, was
applied to analyze the experimental data. Considering that the selection
where f(x) represented the theoretical model for reaction mechanism, x
and use of a reaction model can influence the accuracy of the CR
is a loss in weight, Eɑ denotes activation energy have a unit of kJ/mol. A
method, all the reaction models listed in Table 2 were applied to the
is pre-exponential constant (s−1), R is universal gas constant having to
experimental data. Data from models having highest fitting (coeffi-
value of 8.314 J mol−1 K−1 and T is the temperature in kelvin (K). dx/
cient > 0.99) was selected for determining E values with the CR
dt explains the rate of change of sample weight with time as a function
method.
of k.
dT
( )
When heating rate β = dt °C/min. is considered constant, Eqs. (2)
and (3) can re-written in integrated form as, 2.6. Potentially toxic elements (PTEs) analysis
x dx A T E
g (x ) = ∫0 f (x )
=
β
∫T 0 exp ⎛⎝− RT ⎞ dT
⎠ (4)
The distribution pattern of potentially toxic elements (PTEs) in
bottom ash, fly ash and flue gas was investigated during combustion
Here in Eq. (4) g (x ) is integral function and x is defined as follows, and co-combustion of coal, biomass, biochar, and blends in the fixed-
bed laboratory-scale tube reactor. The temperature of the reactor was
x = (mi−mt )/(mi−mf ) (5)
set 1000 °C during this study. When the temperature was reached to
mi is sample weight at initial stage, mt is sample weight at time t, and 1000 °C, under constant flow of 100 ml air min−1 (∼21.05 ml oxygen
sample weight at final stage is represented by mf in Eq. (5). (O2) min−1 and ∼78.95 ml nitrogen (N2) min−1 as a carrier gas),
When integration of Eq. (4) was performed by taking a natural log ∼5 g of the desired fuel sample having initial partical size of < 50 µm
and applying Coats Redfern equation, it becomes, (coal, biomass, biochar and their blends) in the form of pellet (prepared

g (x ) AR 2RT ⎤ Ea
ln ⎡ 2 ⎤ = ln ⎡ ⎛1− ⎞ − Table 2
⎣ T ⎦ ⎢
⎣ βE ⎝ E ⎠⎥
⎦ RT (6) Equations of models for solid-state reactions.

For activation energy (Ea) and combustion temperature range the Model f(α) G(α)
components; (ln[AR/βE (1 − 2RT/E)] of Eq. (6) becomes constant.
Chemical reaction
Thus, if combustion is assumed to be a 1st-order reaction, then plot of
1 First-order 1−x −ln(1 − x)
ln[g (x )/ T 2] vs. 1/ T provides a straight line with higher correlation 2 Second-order (1 − x)2 (1 − x)−1 − 1
coefficient. The activation energy (Ea) can be computed from line Ea/ R 3 Third-order (1 − x)3 [(1 − x)−2 − 1]/2
slop. Limiting surface reaction between both phases
The thermal degradation of high purity standard material (calcium 4 One dimension 1 x
carbonate (CaCO3), having purity of > 99%) was taken in a 100 ml air 5 Two dimensions 2(1 − x)1/2 1 − (1 − x)1/2
min-1: ∼21.05 ml oxygen (O2) min−1 and ∼78.95 ml nitrogen (N2) 6 Three dimensions 3(1 − x)2/3 1 − (1 − x)1/3
min−1 as a carrier gas, using 5, 10, 15, 20 °C min−1 linear heating Diffusion
program from room temperature to 1000 °C on a thermal analyzer SDT 7 One-way transport 1/2 x x2
8 Two-way transport [−ln(1 − x)]−1 x + (1 − x)ln(1 − x)
Q600 (TA Instruments, USA) equipped with data acquisition system. To
9 Three-way (2/3)(1 − x)2/3/1 − (1 − x)1/ [1 − (1 − x)1/3]2
minimize variability between the experiments, the sample was carefully transport 3

weighted to be between 15 ± 1 mg before degradation testing at the

145
B. Yousaf et al. Applied Energy 208 (2017) 142–157

using laboratory Pellet Press (Model: CIT-LPM-12TB) with Built-in + N)/C and (O + N + S)/C were calculated. The H/C and O/C atomic
Hydraulic Pump LPM-12TB, Columbia International) was taken in a ratios indicated the degree of aromaticity and carbonization. The raw
ultra-pure quartz boat and introduced into the combustion zone of the biomass materials had higher atomic ratios than coal, while pyrolysis
reactor and kept there for 30 min (three replications of each). The products had lower ratios and further lowest atomic ratios of these
bottom-ash samples were collected after each combustion run from the elements were observed in biochars produced at 700 °C. The me-
reactor, ground with pestle-mortar to a particle size < 150 µm and chanism behind the reduction of O/C and H/C molar ratios was might
sieved for further analysis. The fly ash and flue gas samples during each be the dehydration and carbonization reactions, resulting more hy-
run were collected iso-kinetically from the outlet by using a filter drophobic surface of biochars with higher degree of aromaticity de-
system equipped with ultra-fine fiberglass membrane and XAD-2 resin veloped [32]. These results suggested the improved fuel qualities of
sleeve (Fig. 1). These samples were stored below 0 °C to avoid chemical biochars due to a decrease in energy loss, water vapors and smoke
changes before PTEs analysis. PTEs in gas-phase was trapped in aqu- production during the biochar-coal co-combustion process [7]. The
eous solution (10% H2O2 + 6% HNO3), according to USEPA method occurrence of polar functional groups in raw biomass feedstock’s was
29. Inductively coupled plasma spectrometry (ICP-MS, model NexION indicated by higher (O + N)/C ratios of these materials [32]. Whereas
2000 P, PerkinElmer, USA) was used for the PTEs determination from the polarity level tended to decrease in pyrolytic products and least
these recovered samples. The total concentration of PTEs in feed coal, polarity was developed in both biomass-derived biochars produced at a
biomass (peanut shell and wheat straw), biomass-derived biochars, co- higher temperature (700 °C).
combustion residues (bottom-ash and particulate-phase), were de- The moisture content of peanut shell and wheat straw raw materials
termined by following microwave-assisted digestion leach method [31]. decreased from 6.24 ± 0.82 to 1.93 ± 0.08 and 6.95 ± 0.7 to
For microwave digestion, ∼0.25 ± 0.1 g of sample was mixed with 1.52 ± 0.11, respectively after thermal treatment process at 700 °C.
10 ml solution of concentrated acids (HNO3:HF:HCl = 3:1:1) and mi- As, the –OH groups retains water molecule with their absorption sites,
crowaved at 140 °C ± 1. After cooling the samples, these were filtered thus these control samples moisture content. Difference in hydrophilic
with nylon millipore filter of pore size 0.22 µm and analyzed using and hydroxyl (OH) groups regulates absorption variation of biochars.
inductively coupled plasma spectrometry (ICP-MS, model NexION 2000 These absorption sites reduced with the structural decomposition of
P, PerkinElmer, USA) for the determination of PTEs. biomass materials due to thermal treatment results water evaporates.
The following equation was used to find out the final concentration Higher biomass moisture content hinders the combustion, thus with
of PTEs Eq. (7) pyrolysis the product biochar quality as solid biofuels improved [33].
Element conc X dilution factor Volatile matter (VM) contents of biochars produced at 500 °C and
PTEs conc (mg/kg) = 700 °C is approximately four and eight times lower, respectively than
Actual weight of sample (7)
the original biomass materials. Thermal decomposition of PS and WS
The certified reference materials (CRMs) of coal (SRM-1632b), organic components during carbonization caused the decrease in the
wheat straw (RM-8494) were used during analysis to ensure the accu- quantity of VM. It becomes easier to modulate the combustion process
racy of data precision and analytical technique. The recovery rates of using lower VM content fuel [33]. Fixed carbon and ash contents of
PTEs were within range of 95.5–104.5% in the certified standards. The pyrolytic products increased with the rise in pyrolysis temperature;
precision level was ± 5 wt% for analyzed metals. The standardized these were lower in biochars produced at 300 °C and much higher in
solutions of different known concentration were run before each batch products synthesized at 700 °C, but biomass feedstocks have much
of samples. lower contents. This increase in ash content during the thermal process
The inorganic elemental composition of coal, biomass, and biochar in biochars might be due to the concentration of inorganic minerals and
powdered ashes were determined by X-ray fluorescence (XRF) spec- formation of their oxides and carbonates. Higher ash content of bio-
troscopy using ARL QUANTX EDXRF spectrometer (Thermo Fisher chars comparative to raw biomass materials can cause the problem of
Scientific Inc., USA). fouling during combustion, which can be overcome by blending the
coal and biochar in appropriate proportions [18]. Thermal treatment of
3. Results and discussion conocarpus waste, wheat straw, grass, pine wood, coconut fibers and
bambusa balcooa under different pyrolysis conditions, their yields,
3.1. Physico-chemical characteristics of coal, feedstock biomasses, and proximate and ultimate analysis results are comparable to the present
biomass derived-chars study [7,15,32,34].
The concentration of potentially toxic elements (PTEs) in biomass
The assessment of feedstock material basic physico-chemical char- materials and biomass- derived biochars is highly variable and depends
acteristics prior to performing pyrolysis is crucial. The proximate and on the production conditions of biomass, which can be sink or source of
elemental composition of bituminous coal, peanut shell (PS), wheat a particular element [35]. PTEs content in coal, biomass feedstocks, and
straw (WS) and biomass-derived chars is presented in Table 3. The biomass derived biochars synthesized at different pyrolysis tempera-
biomass pyrolysis products had a higher concentration of carbon (C) tures (300, 500 and 700 °C) is shown in Table 4. In a comparison of raw
and lower content of oxygen (O) than the feedstock materials. As the biomass materials, all PTEs content were higher in biochars may be due
pyrolysis temperature increased the fuel quality of synthesized biochars to more loss in mass of the organic fraction during pyrolysis process
upgraded due to a reduction in oxygen and enrichment of carbon than the PTEs mass loss. Thus, enrichment of all PTEs except boron (B)
contents. The amount of nitrogen (N) and sulfur (S) in solid bio-fuel is and bismuth (Bi) in biochar occurred with increases in pyrolysis tem-
of particular concern due to their transformation into gaseous pollu- perature from 300 °C to 700 °C. Whereas, the concentration of B and Bi
tants in the form of NOx (nitrogen oxides) and SOx (sulfur oxides), re- were noticed to higher in the raw biomass materials than biomass-de-
spectively [7]. The concentration of these two elements in feedstock rived biochar. As most PTEs (70–80%) remains in biochar, therefore,
biomass materials was lower than in coal, and their contents in similar enrichment of these noxious pollutants in pyrolytic biochars occurred
pyrolytic products further decreased with increase in the highest with the rise in the thermal treatment temperature. Cu, Cr, Ni and Zn
heating temperature (HHT). Thus, biochar-coal blends combustion as a concentrations were observed higher than other elements in biochars.
potential clean energy source also has additional environmental bene- During pyrolysis process, the partitioning of PTEs in different phases
fits regarding the green house and acid rain forming gases emission (biochar, bio-oil and gaseous products) can happen, depending upon
reduction [7]. These are the most important features of solid fuels used pyrolysis highest heating temperature (HHT) and metal species [27].
in commercial co-firing systems for clean energy production. To study Thermal decomposition of biomass organic components caused the
the variation in atomic composition, molar ratios of H/C, O/C, (O release of bonded PTEs into biochar matrix, resulting major proportion

146
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Table 3
Physico-chemical characteristics of feed coal, biomasses (peanut shell and wheat straw) and biomass-derived biochar fuels.

Parameter Coal Feedstock PS Biochar Feedstock WS Biochar

PSBC300 PSBC500 PSBC700 WSBC300 WSBC500 WSBC700

Lignocellulosic compositional of feedstock (%)


Cellulose – 37.45 ± 0.87 – – – 31.45 ± 1.24 – – –
Hemicellulose – 21.18 ± 0.58 – – – 27.86 ± 0.75 – – –
Lignin – 29.74 ± 0.25 – – – 21.22 ± 0.14 – – –

Proximate and ultimate analysis


C (%) 68.39 ± 1.14 47.88 ± 2.45 61.72 ± 2.12 74.14 ± 2.45 79.11 ± 2.11 39.01 ± 0.95 49.81 ± 1.46 62.81 ± 0.89 70.29 ± 1.73
H (%) 6.09 ± 0.12 6.12 ± 0.42 4.64 ± 0.62 2.72 ± 0.45 1.83 ± 0.45 6.77 ± 0.43 3.99 ± 0.92 2.71 ± 0.11 1.98 ± 0.24
N (%) 0.53 ± 0.08 0.32 ± 0.05 0.28 ± 0.07 0.22 ± 0.09 0.20 ± 0.04 2.14 ± 0.16 1.04 ± 0.25 0.59 ± 0.15 0.45 ± 0.08
S (%) 0.77 ± 0.02 0.46 ± 0.09 0.45 ± 0.08 0.40 ± 0.09 0.36 ± 0.07 0.69 ± 0.09 0.63 ± 0.07 0.55 ± 0.09 0.51 ± 0.05
O (%) 10.87 ± 0.31 43.20 ± 1.65 21.38 ± 0.95 5.58 ± 0.83 2.07 ± 0.26 46.53 ± 1.85 36.39 ± 0.62 22.87 ± 0.65 14.95 ± 0.21
Moisture (%) 2.08 ± 0.12 6.24 ± 0.82 3.54 ± 0.55 2.25 ± 0.12 1.93 ± 0.08 6.95 ± 0.75 2.84 ± 0.12 2.14 ± 0.014 1.52 ± 0.11
VM (%) 33.47 ± 0.84 72.90 ± 1.88 45.12 ± 1.43 15.58 ± 0.95 8.46 ± 0.38 69.06 ± 2.45 35.65 ± 0.86 19.46 ± 0.33 8.67 ± 0.29
Ash (%) 13.35 ± 0.39 2.02 ± 0.14 9.53 ± 0.91 10.94 ± 1.14 12.03 ± 0.25 4.86 ± 0.99 8.15 ± 0.25 9.97 ± 0.18 11.82 ± 0.38
FC (%) 51.10 ± 0.95 18.84 ± 0.72 41.81 ± 1.38 71.23 ± 2.88 77.58 ± 1.85 19.13 ± 0.81 53.36 ± 1.19 68.43 ± 0.95 77.99 ± 0.99
O/C 0.12 ± 0.03 0.68 ± 0.08 0.26 ± 0.08 0.06 ± 0.01 0.02 ± 0.01 0.90 ± 0.05 0.54 ± 0.05 0.27 ± 0.04 0.16 ± 0.03
H/C 1.07 ± 0.09 1.53 ± 0.12 0.90 ± 0.09 0.44 ± 0.05 0.28 ± 0.03 2.08 ± 0.17 0.96 ± 0.12 0.52 ± 0.08 0.34 ± 0.09
(O + N)/C 0.13 ± 0.01 0.68 ± 0.03 0.26 ± 0.04 0.06 ± 0.01 0.02 ± 0.01 0.94 ± 0.08 0.57 ± 0.07 0.29 ± 0.04 0.16 ± 0.07
(O + N+S)/C 0.13 ± 0.01 0.69 ± 0.04 0.27 ± 0.04 0.06 ± 0.01 0.02 ± 0.01 0.95 ± 0.09 0.57 ± 0.09 0.29 ± 0.04 0.17 ± 0.07

HHV (MJ kg−1) 29.85 ± 0.39 18.10 ± 0.27 23.99 ± 0.19 27.85 ± 0.55 28.79 ± 0.34 15.54 ± 0.28 16.82 ± 0.45 21.37 ± 0.26 24.07 ± 0.35

of these elements accumulate in pyrolytic biochar. 3.2. Energy characteristics of feed coal, biomass materials, and coal-
Alkaline-earth metal oxides and alumino-silicate oxides are the biomass blends
basic chemical constituents of PS, WS, and their biochars, dissimilar
from the chemical composition of coal (Table 4). Higher concentration The combustibility of solid fuels can be evaluated by measuring
of these alkaline-earth metal oxides in biomass materials and biomass- their higher heating values (HHV), and HHV values should be higher
derived biochars can be a source of fouling and influence the com- than 20 MJ/kg for a solid fuel to ensure auto-thermal combustion [7].
bustion process [35]. These problems can overcome by blending the The energy content (HHV) of solid fuel (biochar) increased with in-
coal-biochar in appropriate ratios. crease in pyrolysis temperature. The HHV of PSBC700 is comparable to
coal. Thus it can be a potential alternative solid biofuel source. Overall
pyrolysis products of WS at different temperatures have less HHV than
the PS derived-biochars. The destruction and the formation of high-
energy chemical bonds with pyrolysis process considerably increased

Table 4
Potentially toxic elements (µg g−1) and ash minerals composition (%) of feed coal, biomasses (peanut shell and wheat straw) and biomass-derived biochars.

Parameter Coal Feedstock PS Biochar Feedstock WS Biochar

PS300 PS500 PS700 WS300 WS500 WS700

−1
Trace elemental concentration (µg g )
As 2.15 ± 0.11 0.08 ± 0.02 0.24 ± 0.08 0.18 ± 0.12 0.15 ± 0.14 0.13 ± 0.05 0.28 ± 0.09 0.31 ± 0.06 0.30 ± 0.14
B 0.98 ± 0.08 1.23 ± 0.09 0.15 ± 0.05 0.11 ± 0.05 0.24 ± 0.09 0.08 ± 0.02 0.13 ± 0.03 0.09 ± 0.02 0.17 ± 0.07
Ba 1.17 ± 0.07 0.85 ± 0.08 0.92 ± 0.09 0.99 ± 0.13 0.97 ± 0.15 0.18 ± 0.07 0.23 ± 0.05 0.41 ± 0.10 0.29 ± 0.08
Be 2.14 ± 0.13 1.21 ± 0.14 1.19 ± 0.15 1.22 ± 0.09 1.27 ± 0.25 0.14 ± 0.05 0.15 ± 0.05 0.17 ± 0.05 0.16 ± 0.07
Bi 0.47 ± 0.08 0.30 ± 0.08 0.21 ± 0.09 0.14 ± 0.04 0.08 ± 0.02 0.09 ± 0.03 0.11 ± 0.03 0.08 ± 0.03 0.03 ± 0.02
Cd 0.36 ± 0.09 0.05 ± 0.02 0.09 ± 0.03 0.15 ± 0.03 0.12 ± 0.03 0.11 ± 0.01 0.15 ± 0.04 0.18 ± 0.04 0.16 ± 0.08
Co 3.24 ± 0.14 1.37 ± 0.25 1.27 ± 0.17 3.17 ± 0.21 3.58 ± 0.18 0.73 ± 0.08 0.92 ± 0.18 0.98 ± 0.15 1.14 ± 0.13
Cr 19.39 ± 0.38 1.08 ± 0.15 1.39 ± 0.18 1.85 ± 0.18 2.18 ± 0.19 3.11 ± 0.27 3.74 ± 0.35 4.13 ± 0.36 5.27 ± 0.54
Cu 13.69 ± 0.18 6.86 ± 0.49 6.60 ± 0.35 7.75 ± 0.25 7.93 ± 0.34 5.41 ± 0.75 5.59 ± 0.24 6.31 ± 0.75 8.02 ± 0.25
Ga 4.08 ± 0.09 0.87 ± 0.19 1.61 ± 0.23 1.34 ± 0.11 2.62 ± 0.29 0.67 ± 0.21 1.54 ± 0.19 1.77 ± 0.21 2.47 ± 0.19
Ni 24.35 ± 0.85 2.29 ± 0.25 4.25 ± 0.31 4.59 ± 0.25 7.41 ± 0.59 3.79 ± 0.49 2.55 ± 0.24 3.59 ± 0.19 6.56 ± 0.46
Pb 9.39 ± 0.28 0.61 ± 0.09 0.87 ± 0.11 2.40 ± 0.14 1.62 ± 0.23 0.15 ± 0.08 0.31 ± 0.13 1.45 ± 0.13 0.87 ± 0.18
Sb 0.82 ± 0.12 0.27 ± 0.08 1.12 ± 0.09 1.33 ± 0.09 1.13 ± 0.17 0.11 ± 0.03 0.22 ± 0.09 0.30 ± 0.08 0.57 ± 0.08
Sn 1.05 ± 0.19 0.13 ± 0.05 0.25 ± 0.08 0.96 ± 0.07 0.55 ± 0.08 0.05 ± 0.02 0.38 ± 0.09 0.59 ± 0.12 0.45 ± 0.09
V 61.85 ± 1.87 0.14 ± 0.06 0.49 ± 0.09 1.13 ± 0.09 1.53 ± 0.33 0.85 ± 0.12 1.28 ± 0.14 2.14 ± 0.36 2.88 ± 0.14
Zn 67.01 ± 2.15 18.63 ± 0.58 22.27 ± 0.68 31.65 ± 0.88 30.03 ± 1.82 39.28 ± 0.85 44.97 ± 1.18 47.62 ± 0.72 49.39 ± 0.89

Ash composition (%)


Al2O3 37.53 ± 0.74 2.97 ± 0.28 3.28 ± 0.26 3.45 ± 0.18 3.98 ± 0.25 1.38 ± 0.16 1.73 ± 0.19 1.81 ± 0.23 1.95 ± 0.31
CaO 4.18 ± 0.12 19.65 ± 0.39 19.87 ± 0.65 20.15 ± 0.39 20.28 ± 0.63 24.64 ± 0.74 24.92 ± 0.62 24.95 ± 0.35 25.15 ± 0.75
Fe2O3 3.34 ± 0.14 4.94 ± 0.21 5.45 ± 0.34 5.62 ± 0.25 5.85 ± 0.29 3.42 ± 0.21 3.94 ± 0.28 4.25 ± 0.18 3.86 ± 0.32
K2O 1.27 ± 0.09 17.95 ± 0.38 18.08 ± 0.48 18.12 ± 0.46 18.15 ± 0.72 16.33 ± 0.44 16.48 ± 0.35 16.59 ± 0.62 16.64 ± 0.36
MgO 0.95 ± 0.08 11.18 ± 0.29 13.65 ± 0.39 13.24 ± 0.28 12.48 ± 0.55 10.49 ± 0.18 11.63 ± 0.26 12.34 ± 0.42 12.45 ± 0.19
P2O5 0.37 ± 0.08 2.46 ± 0.11 2.53 ± 0.12 2.58 ± 0.15 2.61 ± 0.19 1.76 ± 0.19 1.85 ± 0.13 1.93 ± 0.21 1.99 ± 0.20
SO3 0.19 ± 0.02 1.55 ± 0.17 1.32 ± 0.17 1.15 ± 0.11 0.84 ± 0.09 0.54 ± 0.11 0.37 ± 0.09 0.29 ± 0.08 0.22 ± 0.07
SiO2 48.22 ± 0.39 31.75 ± 0.85 31.87 ± 0.55 32.14 ± 0.28 32.29 ± 0.42 35.25 ± 0.35 35.95 ± 0.24 36.25 ± 0.34 36.41 ± 0.39
TiO2 1.05 ± 0.12 0.48 ± 0.09 0.23 ± 0.09 0.11 ± 0.04 0.07 ± 0.02 0.13 ± 0.04 0.17 ± 0.05 0.08 ± 0.03 0.03 ± 0.01

147
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Fig. 2. The experimental and theoretical calorific values as


Higher Heating Values (HHV) of feed coal, biomasses (peanut
shell, wheat straw), biomass-derived biochars (PSBC300,
PSBC500, PSBC700, WSBC300, WSBC500 and WSBC700) and
their blends representing.

HHV of biochar products [25]. The calorific values (CV) of peanut shell 20 °C/min under an oxidizing atmosphere. The major combustion peaks
(19 MJ/kg) and wheat straw (17 MJ/kg) were much lower than coal of PS were extended between 213–396 °C (stage III) and 397–564 °C
(27 MJ/kg). However, with an increase in pyrolysis temperature the (stage IV) with ∼56% and ∼28% weight loss, respectively. While for
calorific values of synthesized biochars increases with maximum values WS biomass primary combustion zone was in stage II with ∼67%
of PS700 (30 MJ/kg) and WS700 (28 MJ/kg) slightly higher than the weight loss in the temperature range of 169–431 °C and at higher
feed coal CV. Hence, pyrolytic products can be used in co-fired power temperature zone (432–538 °C) in stage IV comparatively less intensity
generation systems in greater proportion as compared to raw biomass peak (∼19% WL) was observed. Combustion and respective weight loss
materials. The calorific values (MJ/kg) of feed coal, raw biomass peaks of raw biomass materials were recorded in different stages mainly
feedstocks, biomass-derived pyrolytic products and their blends are il- due to their different lingo-cellulosic composition. During these stages,
lustrated in Fig. 2. Similarly, an increase in energy values of biomass weight loss is attributed to thermal degradation of cellulose, hemi-
materials was reported by Mohanty [34] and Makwarela [36] with cellulose and lignin components of raw materials. The weight loss at a
pyrolysis and torrefaction, respectively. The fuels co-combustion results lower temperature (< 150 °C) is associated with moisture evaporation
indicated that calorific values of biomass-coal blended fuels were lower and light matter volatilization that is not shown in Table 5. In contrary
than biochar-coal fuels. The WSBC products except for WSBC300 with of feedstock biomass materials, feed coal showed a single board peak
50% content of biochar in fuel during co-combustion have lower CV (450–661 °C) probably due to combustion of volatile components (VCs)
than feed coal, while all other pyrolytic products have calorific values and char simultaneously [17]. Similar observations in literature during
in combustion experiment comparable to or higher than coal CV at both combustion of Alfa, wheat straw (WS), pine wood, wood sawdust, su-
levels of biochar blend in solid fuel. Bada et al. [18] carbonized Bam- garcane and coconut fiber raw materials are reported [7,39]. The
busa balcooa at 380 °C and obtained the calorific value of 28.20 MJ/kg weight loss of biomass materials relatively at the lower temperature
comparable to co-combusted solid fuels energy values of this study. At indicated their higher reactivity, less combustion efficacy and higher
300 °C a sudden increase in energy values of biochar might be due to emission of gaseous and particulate matter pollutants [7].
dehydration and devolatilization reactions and ongoing destruction of The combustion profiles of biochars are dissimilar from the corre-
some components of cellulose and hemicellulose [37]. sponding raw materials Figs. 3 and 4(a and d). The DTG peaks of
From above results, it is concluded that low-temperature pyrolysis pyrolytic PS has been shifted to higher temperature indicating the
products of peanut shell can be used economically in co-fired system combustion zone drifting, and this effect is more noticeable with an
with a higher percentage for clean and environmentally-friendly energy increase in pyrolysis temperature from 300 to 700 °C. According to DTG
production. profile, PBSC300 has broad combustion peaks in the range of 153 °C to
399 °C (stage II) and 400 °C to 521 °C (IV) with a weight loss of ∼43%
3.3. Combustion performance and slagging formation in coal, biomass- and and ∼51%, respectively. In lower combustion zone (stage II) PSBC500
biochar-coal co-combustion systems and PSBC700, approximately 58% and 9% thermal decomposition oc-
curred, respectively. Whereas, at higher temperature stage IV ∼31%
Thermogravimetric (TG) and derivative thermogravimetric (DTG) weight loss of PSBC500 was observed. The thermal decomposition of
thermo-grams of feed coal, biomass materials, biomass-derived bio- PBSC300 and PBSC500 occurred in two stage (II and IV). While
chars and their blends in different ratios at a heating rate of PBSC700 behaves more like coal with major combustion zone (∼76%
20 °C min−1 are depicted in Figs. 3 and 4 represent combustion beha- weight loss) laying in a single zone (stage IV) with a temperature range
vior and thermal stability. The maximum weight loss during thermal of 419–716 °C. As, the combustibility of PSBC700 is comparable to coal,
decomposition of materials corresponding to specific temperature range so it has the promising potential of co-firing in existing coal-fired power
during combustion and co-combustion process is shown in Table 5. plants after blending in higher proportion. In the case of WS pyrolytic
Thermal decomposition process consisted of different stages depending products, the main combustion peaks of WSBC300 have detected in the
upon the combustion temperature; dehydration stage (25–135 °C), temperature range from 247 °C to 373 °C (stage II) with ∼43% weight
combustion and volatile components removal stage (200–500 °C), char reduction and 374–424 °C (stage III) with ∼22% loss in weight. The
oxidation and burnout stage (500–800 °C) [38]. concentrated peaks of WSBC500 and WSBC700 were noticed with a
Thermal decomposition of raw biomass materials (PS and WS) oc- single combustion stage II in the temperature range of 245–477 °C
curred in two distinct main phases when heated at a heating rate of (∼-72%) and 267–507 °C (∼-64%), respectively. Overall pyrolysis has

148
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Fig. 3. The TG (a–c) and DTG (d–e) curves of feed coal, peanut shell, peanut shell-derived biochars (PSBC300, PSBC500 and PSBC700) and their blends at 20 °C min−1.

little impact on total burnout percentage of both biomass materials. The reduction in weight loss as pyrolysis temperature increase was
Pyrolytic biochars have higher combustion efficiency and lower emis- due to intense de-volatilization, less volatile matter (VM) and higher
sions in comparison to PS and WS raw materials. Furthermore, fuel fixed-C (carbon) content of pyrolysis products is responsible for coal
performance is affected by its moisture content. As biomass materials like burning characteristics. These also influenced the ignition tem-
moisture vaporizes during pyrolysis process resulting not only com- perature, pyrolytic products, and raw biomass materials both have
bustion temperature increases but also reduced the emission of carbon lower ignition temperature and found to be more reactive than that of
monoxide (CO) when carbonized products are combusted [7]. coal used in this study [18].

Fig. 4. The TG (a–c) and DTG (d–e) curves of feed coal, wheat straw, wheat straw-derived biochars (WSBC300, WSBC500 and WSBC700) and their blends at 20 °C min−1.

149
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Table 5
Thermochemical characteristics of feed coal, biomasses (peanut shell, wheat straw), biomass-derived biochars and their blends in different steps at 20 °C min−1.

Stage II Stage III Stage IV Burnout (%)

Sample Range (°C) Tmax (°C) DTGmax WL (%) Range (°C) Tmax (°C) DTGmax WL (%) Range (°C) Tmax (°C) DTGmax WL (%)
(%/°C) (%/°C) (%/°C)

C 450–761 624 0.792 87.12 85.97


C80PS20 174–399 348 0.142 11.31 400–748 622 0.577 78.02 88.98
C50PS50 152–407 341 0.345 29.37 408–534 493 0.117 14.67 535–735 610 0.419 44.97 89.16
PS 148–212 192 0.068 5.89 213–396 341 0.676 56.21 397–564 436 0.221 27.94 91.26
C80PSBC300-20 271–369 347 0.057 3.38 370–535 494 0.149 15.87 536–739 618 0.561 70.73 89.64
C50PSBC300-50 331–381 349 0.151 8.36 382–557 521 0.315 39.21 558–734 604 0.398 42.45 90.18
PSBC300 153–399 376 0.246 42.75 400–521 452 0.351 50.61 90.67
C80PSBC500-20 336–479 468 0.119 2.19 479–554 566 0.243 16.75 555–742 626 0.732 67.24 88.86
C50PSBC500-50 342–574 527 0.464 36.65 575–732 621 0.438 53.82 89.75
PSBC500 326–474 451 0.253 58.47 475–566 509 0.327 31.58 88.17
C80PSBC700-20 364–585 578 0.561 29.41 586–738 613 0.673 60.64 87.63
C50PSBC700-50 359–591 572 0.728 54.66 592–729 591 0.466 34.24 88.14
PSBC700 353–418 428 0.159 9.48 419–716 519 0.432 75.62 87.39

C80WS20 177–396 317 0.143 11.51 397–755 615 0.594 74.07 87.93
C50WS50 172–443 312 0.313 30.86 444–751 582 0.416 52.35 88.15
WS 169–431 306 0.621 66.92 432–538 469 0.233 18.64 89.67
C80WSBC300-20 251–467 428 0.119 12.81 468–716 587 0.596 73.56 87.99
C50WSBC300-50 246–464 414 0.267 31.07 465–582 531 0.348 30.64 583–713 604 0.248 19.64 88.16
WSBC300 247–373 338 0.372 42.93 374–424 394 0.261 21.76 425–518 443 0.179 11.16 88.62
C80WSBC500-20 257–507 461 0.184 12.42 508–715 616 0.695 73.57 86.95
C50WSBC500-50 253–491 449 0.478 30.96 492–712 536 0.366 48.34 87.19
WSBC500 245–477 402 1.097 72.12 87.44
C80WSBC700-20 269–522 504 0.304 13.02 523–714 598 0.656 72.01 86.64
C50WSBC700-50 255–521 484 0.468 31.34 522–713 562 0.361 46.87 87.18
WSBC700 267–507 418 0.891 64.02 86.82

Tmax: temperature when weight loss rate is maximum; DTGmax: maximum weight loss rate; WL: weight loss at individual stage.

Peaks at lower temperature are attributed to full volatilization of in stage II and III are associated mainly with the burning of low mo-
cellulose and hemicellulose components [17]. Biomass materials have lecular weight volatile compounds in raw biomass materials, while
lower ignition temperature and Tmax (°C) than the thermally treated stage IV weight loss peaks are primarily linked to coal with a minor
biochar products because the ignition temperature is dependent on the contribution of biomass high molecular weight compounds. The more
release of volatiles and during biomass combustion these releases at an DTGmax of coal-biomass blends containing a higher proportion of coal
early stage [17]. The pyrolyzed rice husk fuel at 400 °C and 500 °C indicates fixed carbon (FC) incomplete combustion. The higher CO
showed one DTG peak at 464 °C and 468 °C, respectively [40]. Simi- concentration and lower CO2 compositions of flue gas and high un-
larly, pinewood pyrolytic biochar combustion occurred in a single burned carbon in fly ash of fuels may represent the combustion beha-
temperature stage (338–499 °C) with a peak at 426 °C [7]. These results viors of fuels (Table S5). Whereas, a greater proportion of biomass in
are dissimilar from present study might be due to different pyrolysis blends reduces the ignition and burnout temperature resulting mini-
conditions, such as HHT, retention time, heating rate, gas flow rate and mized the time and unburnt carbon loss [42]. High weight loss was
biomass particle size, as well as the nature of biomass feedstock ma- found in the biomass-coal blend having greater biomass contents due to
terials used in this study. With the increase in homogeneity and fixed the presence of more volatile components in biomass materials. How-
carbon content of biofuels, their oxidation takes place easily. Thus, with ever, in other hand, the higher content of char in biochar-coal blend
the development of carbon rich, the porous structure of pyrolytic pro- during combustion process caused the better combustion of fuels.
ducts may enhance their reactivity. The shifting of DTG curves to higher The existing co-fired power plants are operating with maximum 10
degrees is associated with oxidation of biochars and improved re- percent biomass in coal-biomass blends. However, simulation based
activity [17,41]. The biochars have improved fuel characteristics with studies indicated that biomass thermal treatment prior to use in the co-
combustion shift to higher temperature and presence of combustion firing system provides practical opportunity to increase the biomass
zones in a continuous manner. percentage in blends [17]. In this work, the pyrolyzed products used for
Biomass proportion is a key factor that influences the co-combustion co-combustion study have promising physicochemical characteristics.
process of coal-biomass blends. When co-combustion of coal-biomass Due to the shift of combustion temperature to a higher level, the
blends were performed, it is evident from DTG thermographs combustion stages of biochar/coal overlap more in comparison of the
(Figs. 3(e) and 4(e)) that blend weight loss peaks lie between the in- respective coal-biomass blends and show better reactivity in compar-
dividual PS and WS fuel combustion zones. Biomass ratio increase in ison of individual fuels. TGA and DTG curves of PS and WS biochar
blends caused the combustion profiles shift towards left with broader blends with coal in the different ratio are illustrated in Figs. 3 and 4(e,
peaks indicating burning takes place at a relatively wider temperature f). The low-temperature pyrolytic product (PSBC300), co-combustion
range in a continuous fashion. When PS and WS raw biomass materials with coal, showed two-stage combustion with higher weight loss rate
have less proportion (20 wt%) in blends, fuels ignition behaves more (DTGmax) in stage IV, and peak separation becomes prominent with an
like coal and major weight loss occurred (∼−75%) in stage IV. Also, increase in weight loss in stage III as the biochar quantity in blended
with an increase in raw biomass materials weight percentage (50%) in fuel increased. Thus, combustion occurred continuously in broad tem-
the blending of fuels, three-stage combustion was identified for perature range with higher biochar percentage in fuel. In the case of
C50PS50, with major weight loss in stage II (∼29%) and stage IV PSBC500, two distinct stage (III and IV) fuel combustion was identified
(∼45%). Whereas, for C50WS50 two-stage weight loss was observed, with the major reactive zone (DTGmax 0.732 %/°C) in stage IV at higher
∼−31% in stage II and stage IV ∼−52%. The combustion zone peaks coal ratio and shifted to lower temperature stage II with ∼37% weight

150
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Fig. 5. Slagging and fouling problems in un-washed and hydro-


thermally washed coal, biomass, biochars and their blends
during combustion.

loss when biochar ratio is 50% in the blend. When the PSBC700-coal fuel overlap with the coal combustion thus pyrolysis upgraded the
blended fuel in equal proportion was combusted its major reactivity biomass materials fuel properties. All peak combustion temperatures
(DTGmax (0.728%/°C) and weight loss stage was observed II. Fur- have been decreases considerably in comparison to coal. Biomass pyr-
thermore, C50WSBC300-50 co-combustion occurred continuously with olytic products aided ignition to coal at low temperature due to mole-
broad range reactivity (II, III, and IV). While WSBC500 and WSBC700 cular structure modifications with pyrolysis higher fixed carbon and
co-combustion behavior is similar; at a lower ratio (20%) major reac- low volatile contents (VCs) [18]. These results suggested that thermo-
tion stage is IV and with higher biochar ratios (50%) fuel ignition in- chemical proprieties of different biomass materials improved with the
creased at a lower temperature (stage II). Overall, the burnout % of fuel pyrolytic route and the resulting biochar can be co-utilized effectively
slightly increases with higher biochar ratio in the blend. in co-firing energy production system with coal in a higher percentage.
All co-combusted samples (80% coal/20% biochar) were burned at The pine wood biochar co-combustion with coal showed similar results
higher temperature showing combustion profiles relatively more com- to our study, but rice husk [7] and wood chips combustion with coal
parable to coal. While samples 50C50BC (50% coal/50% biochar) ex- provide different results [40]. These contradictory results might be due
hibited low temperatures combustion more similar to biochars. Burnout to the heterogeneous nature of blends or may be attributable to dif-
time of fuels increased when coal-biochar blends were combusted and ferent pyrolysis conditions of biomass materials used in this co-com-
DTGmax (weight loss rates) shifted to a lower temperature when biochar bustion experiment. However, slagging formation is important issue
have higher ratios in the mixture. Furthermore, biochar combustion in during co-combustion of biochar-coal due to higher contents of

151
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Fig. 6. CO and CO2 emissions during coal, biomass- and biochar-coal co-
combustion systems.

Fig. 7. Un-burned carbon, soot formation and fly ash yield during coal,
biomass- and biochar-coal co-combustion systems.

inorganics in biochar, which need significant attention. The results of overcome these problems, some pre-treatment should be taken to re-
slagging formation in ash obtained from the coal, biomass, biochar and duce the slagging and fouling problems in biochar-coal co-combustion
their blends are presented in supplementary information (Fig. 5 and system prior to use. The results (Table S6) indicated that the hydro-
Table S6). It can be seen that the addition of raw biomass and pyrolytic thermal pretreatment is effective to reduced slagging and fouling
biochar to coal increased the slagging index and fouling index in blend during co-combustion of biomass-coal or biochar-coal blends.
ashes in comparison to the ash of coal alone, especially for WS and its
biochar. The slagging index and fouling index results confirm that the 3.4. CO and CO2 emissions, un-burned carbon in fly ash and soot yield
slagging and fouling problems encountered during raw biomass and
coal co-combustion were still present when pyrolytic biochars were co- The emission of carbon monoxide (CO) is mainly due to the im-
combusted with coal, especially for agricultural pyrolytic biochars. To perfect oxidation of fuels during combustion. In addition, slow burning

152
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Table 6 (5.41 ± 1015) with biochar produced at 500 °C, however it increased
Kinetic parameters of feed coal, biomasses (peanut shell, wheat straw), biomass-derived again with BC700. Similar trend was observed in the case of soot yield.
biochars and their blends at 20 °C min−1.
Soot formation was higher with PS and PS-based biochar and their
Stage II Stage III Stage IV blends with coal compared to the WS and WS-based biochar and their
Sample E (kJ/mol) R2 E (kJ/ R2 E (kJ/ R2 blends. Trubetskaya et al. [45] studied the effect of biomass type and
mol) mol) origin on formation mechanism and yield of soot. They found that the
soot formed at 1400 °C was more reactive as compared to the soot
C 95.18 0.9975
C80PS20 37.46 0.9975 88.66 0.9934 produced at 1250 °C. In addition they also found that the soot from
C50PS50 45.48 0.9963 32.19 0.9945 65.25 0.9978 almost all biomasses has various alkali contents having nanosized
PS 64.54 0.9935 32.24 0.9936 28.18 0.9951 structure.
C80PSBC300-20 25.34 0.9954 29.68 0.9947 72.37 0.9938
C50PSBC300-50 39.67 0.9973 47.11 0.9964 59.66 0.9965
3.5. Analysis of kinetic parameters
PSBC300 56.25 0.9958 73.76 0.9943
C80PSBC500-20 17.46 0.9976 24.66 0.9953 74.64 0.9973
C50PSBC500-50 57.64 0.9979 53.61 0.9912 Kinetic parameters of feed coal, biomass feedstocks (peanut shell,
PSBC500 51.83 0.9948 66.29 0.9934 wheat straw), biomass-derived biochars and their blends were calcu-
C80PSBC700-20 59.61 0.9957 59.66 0.9962
lated using the Coats-Redfern method and Ist-order reaction mechanism
C50PSBC700-50 74.25 0.9972 51.35 0.9941
PSBC700 39.46 0.9984 79.65 0.9954 presented in Table 6. The higher correlation coefficient (R2) values
(> 0.98) indicated the satisfactory model fitting of experimental data
C80WS20 34.49 0.9938 73.08 0.9968
C50WS50 41.68 0.9985 64.15 0.9975
[38]. Materials reactivity is determined by their activation energy (Ea),
WS 68.85 0.9976 34.82 0.9928 it is the minimum amount of energy required to initiate a reaction and
C80WSBC300-20 28.75 0.9924 67.34 0.9961 considered a potential barrier for most thermo-chemical reactions [10].
C50WSBC300-50 39.05 0.9981 47.22 0.9915 41.38 0.9958 The activation energies of both raw biomass materials are comparable,
WSBC300 52.48 0.9847 38.49 0.9839 30.17 0.9846
but in the meantime, the level of first Ea (stage II) is different with
C80WSBC500-20 42.18 0.9956 76.31 0.9955
C50WSBC500-50 63.22 0.9919 48.54 0.9938 higher values from second energy level (stage IV). The values of raw
WSBC500 102.95 0.9879 biomass materials activation energies in this study are greater than the
C80WSBC700-20 47.65 0.9964 69.45 0.9967 raw Bambusa balcooa and olive tree pruning energy values reported by
C50WSBC700-50 59.33 0.9982 47.39 0.9972 other researchers [17–18]. These differences are due to the diverse
WSBC700 99.74 0.9917
nature of the biomass materials.
E: activation energy (kJ/mol); R2: correlation coefficient. The PS pyrolytic products first activation energy (stage II) reduced
to 39.46 kJ/mol and second Ea (stage IV) increased to 79.65 kJ/mol
of the soot is also another influential phenomena for the formation of when thermal treatment temperature was up to 700 °C. In case of WS
CO, which administrated by diverse kinetics during the last stage of pyrolysis, biochar prepared at 500 °C showed relatively higher Ea
combustion [43]. Fig. 6 demonstrated that the CO formation was higher (102.95 kJ/mol) compared to biochars prepared at 300 °C (52.48 kJ/
with both biomasses and their blends with coal having maximum value mol) and 700 °C (99.74 kJ/mol). The WSBC500 more activation energy
of 203 ± 12 ppm and 182 ± 12 during WS and PS combustion was may be due to lignocellulose reactive components thermal de-
compared with alone coal combustion (167 ± 13), respectively. gradation during pyrolysis. These Ea values were lower than those re-
However, the CO formation during combustion of BC300, BC500 and ported by Yang et al. [26] in pyrolytic products of bamboo. The lower
BC700 and their blends with coal was remarkably decreased with activation energies of peanut and wheat straw derived biochars in the
compared to their raw biomass fuels. This study shows that the use of present study was due to different pyrolysis conditions and composition
biochar produced at high temperature, as solid fuel, have potential to of feedstock materials. In the case of biomass and own biochar-coal
reduce formation and/or emission of CO. In other hand, the increased blends, the Ea values were between individual fuels. However, in
emission of CO with biomass combustion could be attributed to the blends, these values at all combustion stages were lower than the feed
higher lignin contents in biomass and our results (Table 3) related to coal values. The first Ea decreased with a higher ratio of coal, suggested
lignin contents (29.74 ± 0.25 and 21.22 ± 0.14 in PS and WS, re- less quantity of energy is required for initial combustion process of coal-
spectively) confirmed that high lignin contents are responsible for the blended fuels. Whereas, second Ea increased with the higher percentage
formation of CO during combustion of biomass. Another previous study of coal in the blend, pointed out more energy is needed to overcome
suggested that the CO formation was precisely dependent on the com- complex coal structure energy barriers [38]. The trend was opposite
bustion phase and it was also found that the CO formation from the when biomass/biochar percentage was high in fuel; first Ea level ele-
pyrolyzed fuel was much lower than that from any of the other fuels vated and second decreased with higher content. The maximum in-
[44]. Furthermore, the insufficient air can also markedly influence the crease (+40.18 kJ/mol) of first activation energy (stage II) and de-
CO formation through unburned fuel [43]. crease (−27.77 kJ/mol) of second energy level (stage IV) was recorded
The CO2, a main exhaust gas from a combustion system, is an in- when PSBC500 and WSBC500 percentage in blends were increased to
dicator of efficient/complete combustion process. Usually, an efficient 50%, respectively. For the fuel, containing coal-PSBC blend first lowest
combustion increases the CO2 contents in flue gas. Alternatively, in- Ea (stage II) was 17.46 kJ/mol at 20% ratio of biochar synthesized at
efficient/incomplete burning diminishes the CO2, and on the other 500 °C and second lowest Ea (stage IV) was 51.35 kJ/mol with 50%
hand increases CO and hydrocarbons in the form of un-burned carbon ratio of PSBC700. In the case of pyrolytic products of WS, first, lowest
and soot. Fig. 6 illustrates that there were no significant changes in the activation energy was 28.75 kJ/mol with 20% WSBS300 and
CO2 emissions from biomass- and biochar-coal blends and our results C50WSBC300-50 fuel has second lowest Ea (41.38 kJ/mol). The lower
were comparison to the previous study [21]. second activation energy (stage IV) of fuels when biochar percentage in
Fig. 7 illustrates that the value of un-burned carbon content in fly blends were high due to lignin condensation reactions or molecules
ash from WS was maximum (12.3 ± 0.29%), which was markedly escaping tendencies from reaction media as a process carried out at a
higher than coal (10.80 ± 0.35) followed by its blends with coal. higher temperature [25]. Higher temperature or prolonged retention
However, un-burned carbon contents were markedly reduced with time is required for the reactions to complete having higher activation
biochar alone and biochar-coal blends. It was observed that the un- energy (Ea) [10]. The blended fuel activation energies of current ex-
burned carbon contents were decreased to minimum level periment exhibited different trend from the literature may be due to
various conditions used for thermogravimetric analysis, nature of

153
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Fig. 8. The distribution pattern of potentially toxic elements (%) in bottom ash, fly ash and flue gas during combustion of coal, peanut shell, peanut shell-derived biochars (PSBC300,
PSBC500 and PSBC700) and their blends.

pyrolytic products and coal composition [17,42,46]. According to re- PTEs in individual fraction (mg kg−1), Yi indicates the individual yields
sults mentioned above pyrolysis regulate the fuel quality by improving (%) of bottom ash or fly ash. Cf represents the concentrations of PTEs
the combustion and decreasing activation energies of pyrolytic pro- (mg kg−1) in feed fuel.
ducts. Furthermore, the partitioning fraction of volatile PTEs in the flue
gas was calculated using the following mass balance Eq. (9).
3.6. Retention and distribution pattern of PTEs in bottom ash, fly ash and (Cba ·Yba) + (Cfa ·Yfa) ⎤
flue gas during co-combustion process VPTEs = ⎡100−
⎢ Cf ⎥
⎣ ⎦ (9)
The information about the speciation of potentially toxic elements where VPTEs represents the volatile fraction of potentially toxic ele-
(PTEs) during the co-combustion process is pre-requisite to predict the ments (%) in flue gas; Cba and Cfa represent the concentrations of PTEs
performance of combustion system and its detrimental ecological ef- in bottom ash and fly ash (mg kg−1), respectively; Yba and Yfa indicate
fects. The migration and transformation behavior of PTEs (%) in bottom the yields (%) of bottom ash and fly ash, respectively. Cf represents the
ash (BA), fly ash (FA) and flue gas (FG) during combustion of coal, raw concentrations of PTEs (mg kg−1) in feed fuel.
biomass materials (peanut shell, wheat straw), biomass-derived bio- The co-combustion experiment of PS-coal blends indicated that
chars and their blends are depicted in Figs. 8 and 9 and Tables S7 and except Ba, Be, Cd, Ga, and Zn all other PTEs was higher in bottom ash
S8. The distribution of PTEs in different phases depends upon their when biomass proportion in the blend was 50%. The percent increase in
content, interaction between components of blended fuels and with the concentration of As was greatest with a value of +5.10%. Similarly,
minerals as well as combustion temperature in boiler [21]. The parti- during WS-coal blend combustion all PTEs except Ba and Be percentage
tion of PTEs in individual fraction (bottom ash and fly ash) was de- increased in BA as a function of elevated biomass proportion in fuel.
termined using following formula (Eq. (8)). The fly ash (FA) % of Co, Cr, Ni, and V were higher during the burning
Ci ·Yi of fuel containing an equal proportion of raw biomass materials and
IFPTEs = feed coal. However, in flue gas (FG) all PTEs concentration substantially
Cf (8)
reduced when biomass percentage in fuel is higher. The decrease was
where IFPTEs represents the potentially toxic elements (%) of individual maximum for As (−9.04 %), and Cu (−6%) in PS-coal and WS-coal
fraction (bottom ash and fly ash) and Ci represents the concentrations of blended fuels, respectively.

154
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Fig. 9. The distribution pattern of potentially toxic elements (%) in bottom ash, fly ash and flue gas during combustion of coal, wheat straw, wheat straw-derived biochars (WSBC300,
WSBC500 and WSBC700) and their blends.

When the biochars-coal blended fuels retention behavior of PTEs The highest concentration of most studied PTEs was in fly ash
was examined, it was found that in BA percentage of all PTEs excluding compared to BA and FG due to their volatilization and subsequent
Ba and Pb were enriched when biochar ratio in fuel was high condensation reactions on fine particles surface. The Cr, Ni and V
(C50PSBC50). The trend was similar for all three PS pyrolytic products. contents were high in bottom ash indicated their adsorption on rough
However, Ba bottom ash % was slightly reduced during higher pro- surfaces [20]. The flue gas concentration was significantly reduced thus
portion of coal in low-temperature pyrolysis product (C80PSBC300-20). mostly PTEs becomes stabilized and enriched in FA and BA during
The variation in BA concentration of Pb at both levels of all biochars combustion process [35].
blend in fuel was less than 1%, indicated the negligible influence of The volatilization rates of PTEs were higher during combustion of
pyrolysis on retention of Pb. The PTEs enrichment trend in BA of WSBC- raw biomass-coal fuel containing a higher percentage of coal. The As,
coal fuel is similar to PSBC/coal blended fuel except for Ba and Be; the Cd, Pb, Sb, Sn, and Zn have higher volatilization compared to other
retention of these two PTEs showed decreasing trend at higher biochar PTEs, their volatility slightly reduced with higher biomass feedstock
percentage. In fly ash (FA) of biochar-coal blended fuels, the enrich- percentage (50%) in fuel. According to biomass materials ash analysis,
ment of Cr, Co, Ni, and V followed a trend similar to the raw biomass these comprised of more alkaline earth metals than coal. During the
materials. However, Bi, B and Be percentage in FA remain unchanged at combustion process and phase transition reactions of high biomass
both levels of PS biochar in the blend. The Sn augmentation pattern is content fuel, the chemical reactions of alkaline earth metals with sulfide
unique; first, its percentage in FA was more during C50WSBC300-50 and phosphides (Fe3P) might be responsible for the precipitation and
combustion, at a later stage during C50WSBC500-50 and enrichment of PTEs in the bottom or fly ash [21]. PTEs volatilization
C50WSBC700-50 combustion, its concentration depleted. The flue gas percentage during combustion of biochar-coal fuels were lower, and
(FG) emitted percentage of all PTEs excluding Cr in the case of PS effect of decrease volatility pronounced more when biochar percentage
biochar was higher in fuels comprising of 80% coal in the blend, while in fuel was equal to coal (Fig. 10 and Table S9). Morphological study of
Cr content of WS biochar-coal fuel blends at both levels remain ap- the bottom and fly ashes of coal, biomass-coal and biochar-coal was
proximately similar. The transformation behavior of PTEs in different conducted using Scanning Electron Microscope (SEM) to investigate the
phases was somewhat similar to the co-combustion of soybean stalk interaction between biochar and coal particles (Fig. S1 and Fig. 2). The
with coal [20], but diverse from the study of Lynch et al. [2] due to mechanism behind the reduction of PTEs volatility might be due to the
different feedstock materials. increased fixed carbon pool in biochar with pyrolysis, which acted as a

155
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Fig. 10. The level of potentially toxic elements volatilization (%) during co-combustion: (a) volatilization during combustion of coal, peanut shell, peanut shell-derived biochar
(PSBC300, PSBC500 and PSBC700) and their blends; (b) volatilization during combustion of coal, wheat straw, wheat straw-derived biochar (WSBC300, WSBC500 and WSBC700) and
their blends. (n = 3); the tabulated form of this data (including standard deviation: SD ± ) is presented in supplementary information (Table S9).

stable carbon source and promoted PTEs precipitation by providing of these agricultural wastes utility as an energy source as well as eco-
binding sites [47]. The results of the distribution pattern of PTEs in logical benefits. In this study, the fuel properties, kinetics of co-com-
different phases during combustion process are dissimilar from the bustion system, retention and distribution behavior of PTEs in biochar-
previous studies [19,48]. These dissimilarities are due to diverse nature coal blended fuels was assessed. The results showed that energy con-
of biomass materials and biomass pyrolytic products. tents and combustion performance of blended fuels increased sig-
Over all, the major enrichment of all PTEs except Cr, Ni, and V nificantly with pyrolysis of biomass materials. The activation energy of
during combustion and co-combustion experiments were in fly ash pyrolytic products blend with coal was lower than coal alone.
(FA), while these three PTEs concentrations were higher in bottom ash Furthermore, the total PTEs concentration increased in pyrolysis pro-
(BA). However, with pyrolysis, the volatile concentration of all PTEs ducts, but volatilization of these pollutants decreased substantially
decreased considerably in blended fuels at both percentages of biochar. during combustion of coal-biochar blends. Hence, biochars have pro-
PTEs contamination has a serious threat to the environment and hu- mising potential to improve the combustion of low ranked coal and
man’s health during combustion of biomass materials with coal in co- reduce the environmental toxicity risks of PTEs. Furthermore, the data
fired power plants due to the high volatilization potential of some highlighted the influence of pyrolysis temperature on the improvement
elements. Hence, biomass pyrolysis is a good option prior to using in a of biochar combustion quality and reduction of PTEs potential negative
co-firing system to decrease the volatilization of potentially toxic ele- outcomes.
ments.
Acknowledgement
4. Conclusion
The authors greatly acknowledged National Basic Research Program
Pyrolysis of crop wastes (peanut shell and wheat straw) and then co- of China (973 Program, 2014CB238903), National Natural Science
firing of resulting biochars with coal provides an attractive opportunity Foundation of China (41672144, 41402133) and National Key R & D

156
B. Yousaf et al. Applied Energy 208 (2017) 142–157

Program of China (2016YFC0201600) for financial support for this [23] Saikia R, Singh R, Kataki R, Pant KK. Bioresource Technology Perennial grass
(Arundo donax L.) as a feedstock for thermo-chemical conversion to energy and
study. The Chinese Academy of Science (CAS) and The World Academy materials. Biores Technol 2015;188:265–72. http://dx.doi.org/10.1016/j.biortech.
of Science (TWAS) are also greatly acknowledged for providing the 2015.01.089.
[24] Monlau F, Sambusiti C, Antoniou N, Barakat A, Zabaniotou A. A new concept for
CAS-TWAS President fellowship (CAS-TWAS No. 2014-179). We are enhancing energy recovery from agricultural residues by coupling anaerobic di-
also thankful to the China Post-Doctoral Science Foundation. We also gestion and pyrolysis process. Appl Energy 2015;148:32–8. http://dx.doi.org/10.
greatly appreciate the thoughtful comments and valuable suggestions 1016/j.apenergy.2015.03.024.
[25] Yang W, Wang H, Zhang M, Zhu J, Zhou J, Wu S. Fuel properties and combustion
from Dr. Imtiaz M. and Nawaz H.H for the improvement of this kinetics of hydrochar prepared by hydrothermal carbonization of bamboo. Biores
manuscript. Technol 2016;205:199–204. http://dx.doi.org/10.1016/j.biortech.2016.01.068.
[26] Zhu M, Zhang Z, Zhang Y, Liu P, Zhang D. An experimental investigation into the
ignition and combustion characteristics of single droplets of biochar water slurry
Appendix A. Supplementary materials fuels in air. Appl Energy 2017;185:2160–7. http://dx.doi.org/10.1016/j.apenergy.
2015.11.087.
[27] Devi P, Saroha AK. Risk analysis of pyrolyzed biochar made from paper mill effluent
Supplementary data associated with this article can be found, in the treatment plant sludge for bioavailability and eco-toxicity of heavy metals. Biores
online version, at http://dx.doi.org/10.1016/j.apenergy.2017.10.059. Technol 2014;162:308–15. http://dx.doi.org/10.1016/j.biortech.2014.03.093.
[28] Debela F, Thring RW, Arocena JM. Immobilization of heavy metals by co-pyrolysis
of contaminated soil with woody biomass. Water Air Soil Pollut 2012;223:1161–70.
References http://dx.doi.org/10.1007/s11270-011-0934-2.
[29] Shen DK, Gu S, Luo KH, Bridgwater AV, Fang MX. Kinetic study on thermal de-
composition of woods in oxidative environment. Fuel 2009;88:1024–30. http://dx.
[1] Huang YF, Cheng PH, Chiueh P Te, Lo SL. Leucaena biochar produced by microwave
doi.org/10.1016/j.fuel.2008.10.034.
torrefaction: fuel properties and energy efficiency. Appl Energy 2016. http://dx.doi.
[30] Niu Z, Liu G, Yin H, Zhou C, Wu D, Yousaf B, et al. In-situ FTIR study of reaction
org/10.1016/j.apenergy.2017.03.007.
mechanism and chemical kinetics of a Xundian lignite during non-isothermal low
[2] Lynch D, Low F, Henihan AM, Garcia A, Kwapinski W, Zhang L, et al. Behavior of
temperature pyrolysis. Energy Convers Manage 2016;124:180–8. http://dx.doi.org/
heavy metals during fluidized bed combustion of poultry litter. Energy Fuels
10.1016/j.enconman.2016.07.019.
2014;28:5158–66. http://dx.doi.org/10.1021/ef500981k.
[31] Link DD, Walter PJ, Kingston HM. Development and validation of the new EPA
[3] Nzihou A, Stanmore B. The fate of heavy metals during combustion and gasification
microwave-assisted leach method 3051A. Environ Sci Technol 1998;32:3628–32.
of contaminated biomass-A brief review. J Hazard Mater 2013;256–257:56–66.
http://dx.doi.org/10.1021/es980559n.
http://dx.doi.org/10.1016/j.jhazmat.2013.02.050.
[32] Al-Wabel MI, Al-Omran A, El-Naggar AH, Nadeem M, Usman ARA. Pyrolysis tem-
[4] Yousaf B, Amina Liu G, Wang R, Imtiaz M, Shahid Rizwan M, et al. The importance
perature induced changes in characteristics and chemical composition of biochar
of evaluating metal exposure and predicting human health risks in urbaneperiurban
produced from conocarpus wastes. Biores Technol 2013;131:374–9. http://dx.doi.
environments influenced by emerging industry. Chemosphere 2016;150:79–89.
org/10.1016/j.biortech.2012.12.165.
http://dx.doi.org/10.1016/j.chemosphere.2016.02.007.
[33] Liu Z, Fei B, Jiang Z, Liu X. Combustion characteristics of bamboo-biochars. Biores
[5] Lestander Ta, Rudolfsson M, Pommer L, Nordin A. NIR provides excellent predic-
Technol 2014;167:94–9. http://dx.doi.org/10.1016/j.biortech.2014.05.023.
tions of properties of biocoal from torrefaction and pyrolysis of biomass. Green
[34] Mohanty P, Nanda S, Pant KK, Naik S, Kozinski JA, Dalai AK. Evaluation of the
Chem 2014;16:4906–13. http://dx.doi.org/10.1039/C3GC42479K.
physiochemical development of biochars obtained from pyrolysis of wheat straw,
[6] U.S. Energy Information Administration. International Energy Outlook 2016. vol.
timothy grass and pinewood: Effects of heating rate. J Anal Appl Pyrol
0484(2016) doi: www.eia.gov/forecasts/ieo/pdf/0484(2016).pdf; 2016.
2013;104:485–93. http://dx.doi.org/10.1016/j.jaap.2013.05.022.
[7] Liu Z, Balasubramanian R. A comparison of thermal behaviors of raw biomass,
[35] Vassilev SV, Vassileva CG, Baxter D. Trace element concentrations and associations
pyrolytic biochar and their blends with lignite. Biores Technol 2013;146:371–8.
in some biomass ashes. Fuel 2014;129:292–313. http://dx.doi.org/10.1016/j.fuel.
http://dx.doi.org/10.1016/j.biortech.2013.07.072.
2014.04.001.
[8] Tsai WT, Liu SC, Hsieh CH. Preparation and fuel properties of biochars from the
[36] Makwarela MO, Bada SO, Falcon RMS. Co-firing combustion characteristics of
pyrolysis of exhausted coffee residue. J Anal Appl Pyrolysis 2012;93:63–7. http://
different ages of Bambusa balcooa relative to a high ash coal. Renew Energy
dx.doi.org/10.1016/j.jaap.2011.09.010.
2017;105:656–64. http://dx.doi.org/10.1016/j.renene.2016.12.059.
[9] Wang X, Hu M, Hu W, Chen Z, Liu S, Hu Z, et al. Thermogravimetric kinetic study of
[37] Park SW, Jang CH, Baek KR, Yang JK. Torrefaction and low-temperature carboni-
agricultural residue biomass pyrolysis based on combined kinetics. Biores Technol
zation of woody biomass: Evaluation of fuel characteristics of the products. Energy
2016;219:510–20. http://dx.doi.org/10.1016/j.biortech.2016.07.136.
2012;45:676–85. http://dx.doi.org/10.1016/j.energy.2012.07.024.
[10] Gil MV, Casal D, Pevida C, Pis JJ, Rubiera F. Thermal behaviour and kinetics of
[38] He C, Wang K, Yang Y, Wang J. Utilization of sewage-sludge-derived hydrochars
coal/biomass blends during co-combustion. Biores Technol 2010;101:5601–8.
toward efficient cocombustion with different-rank coals: effects of subcritical water
http://dx.doi.org/10.1016/j.biortech.2010.02.008.
conversion and blending scenarios. Energy Fuels 2014;28:6140–50.
[11] Wang G, Zhang J, Shao J, Liu Z, Zhang G, Xu T, et al. Thermal behavior and kinetic
[39] Boumanchar I, Chhiti Y, M’hamdi Alaoui FE, El Ouinani A, Sahibed-Dine A, Bentiss
analysis of co-combustion of waste biomass/low rank coal blends. Energy Convers
F, et al. Effect of materials mixture on the higher heating value: Case of biomass,
Manage 2016;124:414–26. http://dx.doi.org/10.1016/j.enconman.2016.07.045.
biochar and municipal solid waste. Waste Manage 2016. http://dx.doi.org/10.
[12] Arvelakis S, Frandsen FJ. Rheology of fly ashes from coal and biomass co-com-
1016/j.wasman.2016.11.012.
bustion. Fuel 2010;89:3132–40. http://dx.doi.org/10.1016/j.fuel.2010.04.019.
[40] Park SW, Jang CH. Effects of pyrolysis temperature on changes in fuel character-
[13] Celaya AM, Lade AT, Goldfarb JL. Co-combustion of brewer’s spent grains and
istics of biomass char. Energy 2012;39:187–95. http://dx.doi.org/10.1016/j.
Illinois No. 6 coal: Impact of blend ratio on pyrolysis and oxidation behavior. Fuel
energy.2012.01.031.
Process Technol 2015;129:39–51. http://dx.doi.org/10.1016/j.fuproc.2014.08.
[41] Sahu SG, Sarkar P, Chakraborty N, Adak AK. Thermogravimetric assessment of
004.
combustion characteristics of blends of a coal with different biomass chars. Fuel
[14] Liu Z, Quek A, Kent Hoekman S, Srinivasan MP, Balasubramanian R.
Process Technol 2010;91:369–78. http://dx.doi.org/10.1016/j.fuproc.2009.12.
Thermogravimetric investigation of hydrochar-lignite co-combustion. Biores
001.
Technol 2012;123:646–52. http://dx.doi.org/10.1016/j.biortech.2012.06.063.
[42] Li H, Xia S, Ma P. Thermogravimetric investigation of the co-combustion between
[15] Khan AA, de Jong W, Jansens PJ, Spliethoff H. Biomass combustion in fluidized bed
the pyrolysis oil distillation residue and lignite. Biores Technol 2016;218:615–22.
boilers: Potential problems and remedies. Fuel Process Technol 2009;90:21–50.
http://dx.doi.org/10.1016/j.biortech.2016.06.104.
http://dx.doi.org/10.1016/j.fuproc.2008.07.012.
[43] Ithnin AM, Ahmad MA, Bakar MAA, Rajoo S, Yahya WJ. Combustion performance
[16] Liu Z, Balasubramanian R. Upgrading of waste biomass by hydrothermal carboni-
and emission analysis of diesel engine fuelled with water-in-diesel emulsion fuel
zation (HTC) and low temperature pyrolysis (LTP): A comparative evaluation. Appl
made from low-grade diesel fuel. Energy Convers Manage 2015;90:375–82. http://
Energy 2014;114:857–64. http://dx.doi.org/10.1016/j.apenergy.2013.06.027.
dx.doi.org/10.1016/j.enconman.2014.11.025.
[17] Toptas A, Yildirim Y, Duman G, Yanik J. Combustion behavior of different kinds of
[44] Mitchell EJS, Lea-Langton AR, Jones JM, Williams A, Layden P, Johnson R. The
torrefied biomass and their blends with lignite. Biores Technol 2015;177:328–36.
impact of fuel properties on the emissions from the combustion of biomass and
http://dx.doi.org/10.1016/j.biortech.2014.11.072.
other solid fuels in a fixed bed domestic stove. Fuel Process Technol
[18] Bada SO, Falcon RMS, Falcon LM. Characterization and co-firing potential of a high
2016;142:115–23. http://dx.doi.org/10.1016/j.fuproc.2015.09.031.
ash coal with Bambusa balcooa. Fuel 2015;151:130–8. http://dx.doi.org/10.1016/
[45] Trubetskaya A, Jensen PA, Jensen AD, Garcia Llamas AD, Umeki K, Gardini D, et al.
j.fuel.2015.01.068.
Effects of several types of biomass fuels on the yield, nanostructure and reactivity of
[19] Wiinikka H, Grönberg C, Boman C. Emissions of heavy metals during fixed-bed
soot from fast pyrolysis at high temperatures. Appl Energy 2016;171:468–82.
combustion of six biomass fuels. Energy Fuels 2013;27:1073–80. http://dx.doi.org/
http://dx.doi.org/10.1016/j.apenergy.2016.02.127.
10.1021/ef3011146.
[46] Fan Y, Yu Z, Fang S, Lin Y, Lin Y, Liao Y, et al. Investigation on the co-combustion of
[20] Zhou C, Liu G, Xu Z, Sun H, Lam PKS. The retention mechanism, transformation
oil shale and municipal solid waste by using thermogravimetric analysis. Energy
behavior and environmental implication of trace element during co-combustion
Convers Manage 2016;117:367–74. http://dx.doi.org/10.1016/j.enconman.2016.
coal gangue with soybean stalk. Fuel 2017;189:32–8. http://dx.doi.org/10.1016/j.
03.045.
fuel.2016.10.093.
[47] Li H, Ye X, Geng Z, Zhou H, Guo X, Zhang Y, et al. The influence of biochar type on
[21] Zhou C, Liu G, Fang T, Lam PKS. Investigation on thermal and trace element
long-term stabilization for Cd and Cu in contaminated paddy soils. J Hazard Mater
characteristics during co-combustion biomass with coal gangue. Biores Technol
2016;304:40–8. http://dx.doi.org/10.1016/j.jhazmat.2015.10.048.
2015;175:454–62. http://dx.doi.org/10.1016/j.biortech.2014.10.129.
[48] Dong J, Chi Y, Tang Y, Ni M, Nzihou A, Weiss-Hortala E, et al. Partitioning of heavy
[22] Libra Ja, Ro KS, Kammann C, Funke A, Berge ND, Neubauer Y, et al. Hydrothermal
metals in municipal solid waste pyrolysis, gasification, and incineration. Energy
carbonization of biomass residuals: a comparative review of the chemistry, pro-
Fuels 2015;29:7516–25. http://dx.doi.org/10.1021/acs.energyfuels.5b01918.
cesses and applications of wet and dry pyrolysis. Biofuels 2011;2:71–106. http://dx.
doi.org/10.4155/bfs.10.81.

157

Potrebbero piacerti anche