Anda di halaman 1dari 10

Journal of Analytical and Applied Pyrolysis 137 (2019) 96–105

Contents lists available at ScienceDirect

Journal of Analytical and Applied Pyrolysis


journal homepage: www.elsevier.com/locate/jaap

Solid waste to biofuels and heterogeneous sorbents via pyrolysis of wheat T


straw in the presence of fly ash as an in situ catalyst
Lihui Gaoa,b,c, , Jillian L. Goldfarbb,c,d

a
School of Chemical Engineering and Technology, China University of Mining and Technology, Xuzhou, 221116, People’s Republic of China
b
Department of Mechanical Engineering, Boston University, 110 Cummington Mall, Boston, MA, 02215, United States
c
The Leone Family Department of Energy & Mineral Engineering, The EMS Energy Institute, and The Institutes of Energy and the Environment, The Pennsylvania State
University, University Park, PA, 16802, United States
d
Department of Biological and Environmental Engineering, Cornell University, 226 Riley-Robb Hall, Ithaca, NY, 14853, United States

ARTICLE INFO ABSTRACT

Keywords: Solid waste disposal imposes worldwide environmental and economic hardships. To alleviate such issues, the
Wheat straw present work uses coal fly ash to in situ catalyze the pyrolysis of wheat straw, an abundant agricultural biomass,
Coal fly ash and to simultaneously produce heterogeneous sorbents for water treatment. Coal fly ash mixed in varying
Pyrolysis proportions with wheat straw shows a catalytic effect by lowering the temperature at which many non-
Biofuels
condensable gaseous species are released. As the weight fraction of fly ash in the wheat straw increased from 1 to
Heterogeneous biochar
10 wt%, the overall conversion of biomass increased. 1–5 wt% fly ash favored conversion to the gas phase
Catalyst
(especially CO2, CH4 and C2H4), whereas the 10 wt% mixture produced more condensable species with fewer
detectable oxygenated components and increased furan concentration. For a more complete waste-to-byproduct
transformation, the resulting heterogeneous wheat straw-fly ash biochars were used to removed methylene blue,
a model organic pollutant, from water, at higher capacities and faster rates than biochar alone. Overall, the coal
fly ash may be a potential inexpensive catalyst for in situ upgrading biomass pyrolysis, simultaneously producing
heterogeneous sorbents with enhanced adsorption capacities for organic pollutants.

1. Introduction The literature contains many examples of biomass conversion pro-


cesses that use a downstream catalyst to up-convert evolving vapors
Given our limited global fossil fuel supply, increasing energy con- [8,9], and others that focus on in situ processes, whereby the catalyst is
sumption, and the anthropogenic impacts of energy use, considerable added directly to the biomass [10,11]. In both approaches, natural and
attention is focused on the development of renewable fuels [1]. Biomass synthetic aluminosilicates and solid acid catalysts can promote het-
is touted as a clean fuel feedstock as it has a negligible sulfur content, erogeneous biomass conversion reactions to obtain deoxygenated che-
yielding lower SO2 emissions than conventional fossil fuels and has po- micals and fuels [12,13]. Among solid acid catalysts, montmorillonite
tentially net negative CO2 emissions [2]. Among biomass to energy K10 exhibits catalytic activity for cellulose to levoglucosan pyrolytic
conversion technologies, thermochemical techniques such as pyrolysis conversion [14], potentially due to montmorillonite’s dual Brønsted
(heating in the absence of oxygen) offer the potential to produce liquid and Lewis acid sites [15]; Brønsted acidity decreases with a corre-
fuels that could replace fossil fuels for power generation, in transporta- sponding increase in Lewis acidity as temperature increases [16]. Re-
tion applications, and could be used to produce valuable chemicals [3]. cently, our group demonstrated that bentonite clay can improve pyr-
However, pyrolysis bio-oils have high viscosity and acidity, and high olysis gas yield and bio-oil quality via in situ catalytic upgrading [17].
water, oxygen and ash contents, all of which lead to unstable and cor- To improve the economic viability of pyrolytic biomass to biofuel
rosive fuels [4]. Current upgrading processes for pyrolysis bio-oils in- conversions, we must identify low-cost catalysts, including materials
clude hydrodeoxygenation, catalytic cracking, emulsification and steam that need not be recovered and recycled, such as coal fly ash.
reforming [5], but the poor quality, low yield and cost of upgrading Fly ash is produced by burning coal in power stations and other
hampers the commercial application of pyrolysis bio-fuels [6,7]. industrial sources [18]. Every year over 100 million tons of fly ash,


Corresponding author at: School of Chemical Engineering and Technology, China University of Mining and Technology, Xuzhou, 221116, People’s Republic of
China.
E-mail address: gaolihui0709@126.com (L. Gao).

https://doi.org/10.1016/j.jaap.2018.11.014
Received 18 August 2018; Received in revised form 6 November 2018; Accepted 10 November 2018
Available online 15 November 2018
0165-2370/ © 2018 Elsevier B.V. All rights reserved.
L. Gao, J.L. Goldfarb Journal of Analytical and Applied Pyrolysis 137 (2019) 96–105

bottom ash, slag, and flue gas desulfurization products are generated, Table 1
much of which reside in containment ponds that pose environmental Proximate and surface area analysis of raw wheat straw and coal fly ash.
and health risks [19]. Coal fly ash is comprised of 60% to 80% clay Proximate Analysis Wheat Straw (Raw Biomass) Coal Fly Ash Average Wt %
minerals, with kaolinite, quartz and pyrite making up the majority [20]. Average Wt % (dry basis) (dry basis)
SEM-EDS shows that fly ash consists mainly of amorphous alumino-
silicate spheres and a lesser amount of iron-rich spheres [18]. As such, Volatile Matter 86.09 ± 0.79 11.00 ± 0.22
Fixed Carbon 9.27 ± 0.89 12.89 ± 0.06
coal fly ash contains natural catalysts such as iron oxide, lime, alkalis, Ash (Inorganic 4.64 ± 0.11 76.11 ± 0.20
alumina and has been suggested as a possible catalyst for carbonaceous Matter)
wastes such as biomass and plastics to biofuel conversions [19,21–23]. Ultimate Analysis
The present work investigates the suitability of fly ash as an in situ C 53.94 ± 0.56
H 5.93 ± 0.01
catalyst for the pyrolysis of wheat straw. Wheat straw was chosen as a
N 0.70 ± 0.09
target biomass as it is an abundant, inexpensive agricultural waste [24] O 39.32 ± 0.63
that requires significant pre-treatment to be combusted as a solid fuel to S 0.11 ± 0.01
remove high potassium and alkali contents that lead to slagging and
fouling [25]. However, wheat straw pyrolysis is potentially catalyzed Surface Area Analysis Wheat Straw (Raw Biomass) Coal Fly Ash
by the presence of such alkalis. To date, literature reports of wheat
SBET, m2/g 4.752 ± 0.202 30.146 ± 0.527
straw pyrolysis oils show mainly C2 to C10 compounds [26] with a large Vtotal,cm3 0.0095 ± 0.0009 0.0427 ± 0.0016
fraction of oxygenated and aromatic compounds, and at higher heating Vmeso,cm3 0.0081 ± 0.0008 0.0362 ± 0.0014
rates ketones and alcohols (including phenols) [27]. Wheat straw pyr-
SEM/EDS Wheat Straw Ash Average Coal Fly Ash Average Wt %
olysis could be further improved using coal fly ash in the biomass as an
Elemental Wt %
in situ catalyst. By using a slow pyrolysis technique (10 °C/min) we can Analysis
simultaneously produce a heterogeneous biochar sorbent material that
can remove a model organic pollutant from water, potentially con- Al 2.93 ± 3.18
verting these wastes to three sustainable products: pyrolysis gas, bio-oil, Ba 1.65 ± 0.08
Br 0.42 ± 0.07
and a water treatment material. C 31.94 ± 20.16
Ca 1.17 ± 0.01 1.38 ± 3.01
2. Experimental Cl 0.05 ± 0.05
Cr 5.02 ± 0.12
Cu 2.85 ± 0.04
2.1. Samples
Fe 9.35 ± 0.29 1.49 ± 2.64
K 2.82 ± 0.02 1.54 ± 3.65
Wheat straw (Thunder Acres, Kansas, USA) was ground in a coffee Mg 0.46 ± 0.77
mill and sieved to a particle size between 100 and 300 μm to limit Mn 0.05 ± 0.01
transport limitations by maintaining a Biot number less than 0.1 [28]. Na 4.94 ± 7.69
O 30.67 ± 13.53
Fly ash was obtained from a local utility company that uses a blend of P 0.49 ± 0.03
high-volatile bituminous coals with average higher heating values S 6.43 ± 0.65 2.36 ± 3.39
ranging from 22,000–28,000 kJ/kg in two 50 MW boilers. Particles Sc 2.78 ± 0.31
were all less than 150 μm in diameter, with the majority below 70 μm. Si 44.82 ± 2.31 8.48 ± 0.08
Sr 1.50 ± 0.01
Blends of the biomass and fly ash were measured on a semi-micro-
Ti 15.47 ± 0.54 12.39 ± 10.05
balance to the 0.1 mg, then vortex mixed to insure homogeneity. The Zn 2.31 ± 0.31
following abbreviation system denotes these mixtures throughout the V 1.70 ± 0.14
text: WS/FA(a:b) where WS is wheat straw, FA is fly ash, (a:b) is the
mass ratio (w/w).
2.2. Pyrolysis of biomass and biomass-coal ash blends
Proximate analysis was performed by thermogravimetric analysis
using a Mettler Toledo TGA-DSC1. Samples (raw biomass and later bio-
The TGA-DSC was used to determine the biochar yield and pyrolytic
char) were heated in the TGA under N2 to 110 °C at 50 °C/min and held
thermal decomposition behavior using approximately 10 mg samples,
30 min (to establish a dry baseline), then to 900 °C, held for 30 min
heated at 5 and 25 °C/min in high purity nitrogen flowing at 50 mL/min
under N2 (loss due to volatile matter), and finally to 950 °C at 50 °C/min
from 25 °C to 110 °C and held for 30 min to remove moisture, then
under air and held for 30 min (loss due to fixed carbon); residual mass is
heated from 110 °C to 800 °C under N2 at β and held for 30 min to
considered to be inorganic matter, loosely termed “ash.” Results are
exhaust all volatiles. A baseline run at each heating rate was used to
presented as the average of at least three runs with the corresponding
correct for buoyancy.
standard deviation. Ultimate analysis of the wheat straw was performed
To analyze the biofuel produced and the nature of the resulting
on a LECO 628 analyzer equipped with sulfur module for CHN (ac-
biochars, the raw biomass and biomass-coal fly ash blends (0.5-0.6 g)
cording to ASTM D-5373) and S (ASTM D-1552), with oxygen de-
were placed in a porcelain boat in a 2-MTI tube furnace as previously
termined by difference. A semi-quantitative elemental distribution of
described [29]. To determine an appropriate temperature for pyrolysis
the coal fly ash and wheat straw ash was performed via SEM-EDS on a
to compare the impact of blend ratio on biofuel quality, the WS/
Thermo Fisher (FEI) Q250 with gold sputter-coated samples; elemental
FA(20:1) mixture was heated at 10 °C/min in 120 mL/min of flowing
results are shown as the weight and atomic range and average com-
high purity nitrogen to a series of temperatures (450 °C, 550 °C, 650 °C,
position of 12 samples for coal ash and 5 samples for wheat straw ash to
750 °C) and held for 1 h. After analysis of these experiments, 650 °C was
explore the range of elemental compositions in the samples.
chosen as the pyrolysis temperature, and all blends were subsequently
The porous nature of the samples was determined using nitrogen
pyrolyzed at 650 °C at least twice each sample; results (including yields)
adsorption isotherms with a BELSORP-max version 2.1. Approximately
are averaged and if runs deviated by more than 5% they were repeated.
400 mg of each sample was degassed at 80 °C for 10 h under vacuum to
During pyrolysis in the tube furnace, the devolatilized gases were
remove any surface impurities. The surface area was measured via N2
measured online by a Quadrupole Mass Spectrometer Residual Gas
adsorption at 77 K via BET analysis using 11 data points over a partial
Analyzer (Extorr RGA XT300 M) monitoring m/z signals of 2, 16, 26,
pressure range of 0.05-0.35. Proximate analyses and surface areas of
27, 30 and 44, respectively (H2, CH4, C2H2, C2H4, C2H6, CO2) as a
raw materials are given in Table 1.

97
L. Gao, J.L. Goldfarb Journal of Analytical and Applied Pyrolysis 137 (2019) 96–105

Fig. 1. Derivative thermogravimetric curves for wheat straw and wheat straw + fly ash blends.

function of partial pressure using a Pirani gauge (for more information of 0–15 mg/L to an R2 value of 0.99995 (where Beer-Lambert law ap-
on monitoring pyrolysis gas compounds, we refer the reader to Evans plies for this compound). All experiments were performed in duplicate
and Milne [30], and Huang et al. [31]). Gas yield is quantified knowing and the average values are reported.
the total flow rate (controlled by an Omega Mass Flow Controller, ± The adsorption equilibrium experimental data were fit to Langmuir
0.2 mL/min) and pressure from the Pirani gauge assuming the ideal gas (Eq. (1)), and Freundlich (Eq. (2)) isotherms to relate Ce and qe, the
law applies. Condensable gases were trapped in 4 mL of di- equilibrium concentrations of adsorbate in solution and on adsorbent
chloromethane at 4 °C (Acros Organics, HPLC grade). After pyrolysis, surface, respectively.
the solid samples were cooled to room temperature under nitrogen to qmKLCe
prevent oxidation. Langmuir : qe =
The condensable bio-oil components were analyzed on an Agilent
1 + KLCe (1)
gas chromatograph (7890B) - mass spectrometer (5977 A) (GC–MS) in
Freundlich: qe = KF Ce1/ n (2)
split mode with a split ratio of 30:1. 1 μL samples were injected at
250 °C using helium (ultra-high purity,) as the carrier gas at a total flow qm and KL are the maximum adsorption capacity and equilibrium
rate of 1 mL/min. The GC oven was initially set at 45 °C with a hold constant of adsorption associated with complete monolayer coverage;
time of 10 min, followed by heating at 3 °C/min to 250 °C and then held KF indicates adsorption capacity, with 1/n representing the hetero-
for 5 min. The interface temperature was set to 325 °C. Mass spectra geneity factor. Data from kinetic experiments were fit to pseudo-first
were recorded under electron ionization mode within the m/z range of and second-order models to assess the mechanism(s) controlling the
29–300 from 7 to 80 min, with a 7-minute solvent delay. The GC–MS overall rate of adsorption and determine rate constants (k) as a function
was calibrated using biofuel “marker” compounds, which were chosen of qe and adsorption at time t, qt.
as they are detected in biofuels from a variety of biomasses across the
literature. Each compound was calibrated with over 7 points between dq
Pseudo first order: = k1 (qe qt )
10 and 700 ppm, with a minimum R2 value of 0.996. Additionally, the dt (3)
chromatograph was semi-quantitatively analyzed by integrating the top
dq
(by area) 20 gas chromatogram peaks. Peaks are reported with relative Pseudo second order : = k2 (qe qt ) 2
area to the nearest calibrated compound if their NIST-library identifi-
dt (4)
cation similarity was greater than 90% and their identity compared to
prior compounds detected in the literature. Samples were analyzed at 3. Results and discussion
least twice; errors were estimated by frequent (daily) injections of
known mixtures and are less than 5 ppm for every compound. Table 1 details the characteristics of the raw biomass and wheat
Compounds that were calibrated for are reported in Tables and Figures straw. There is a considerable amount of residual pyrolyzable matter
as mgcompound/gbiomass; those that were semi-quantitatively analyzed via (11 wt%) left in the coal fly ash. Elemental analysis shows the fly ash is
chromatogram area are reported as total ion count (TIC)/gbiomass. a heterogeneous mixture of potentially catalytic materials such as Na,
Mg, Mn, etc. as is often observed throughout the literature [18]. As
2.3. Evaluation of biochars’ sorption behavior shown in the DTG curves of Fig. 1, there was minimal difference ob-
served in the peak devolatilization temperatures for the wheat straw
To determine the impact of co-pyrolysis on the relative adsorption (WS) and WS + fly ash (FA) blends. As suggested by ICTAC Kinetics
capacities and uptake rates of each biochar, we ran methylene blue Committee recommendations for thermal analyses, the peak reaction
(MB) adsorption experiments. Batch adsorption experiments used a 1: rate scales with heating ramp rate (a difference of 5 times faster de-
400 biochar : solution ratio at room temperature (25.5–27.0 °C). The composition at peak temperature, which shifts from ∼ 590 K to 615 K
adsorption isotherms were performed with MB solution concentrations in going from 5 to 25 °C/min), as can be seen in the y-axis scales of
of: 20, 50, 80, 100, 150, 200, 250 mg/L and allowed to equilibrate for Fig. 1 [32]. This suggests minimal change in the overall rate of con-
48 h. Adsorption kinetics were determined using an 80 mg/L MB solu- version of the pure versus “catalyzed” biomass. Such a finding is not
tion and sampled at time intervals of 1, 2, 5, 10, 20, 30 min, and 1, 2, 4, uncommon; Nowakowski et al. found a decrease in volatiles produced
6, 8, 24 h. Samples were centrifuged for 1 min and the dye concentra- for K-impregnated short rotation willow coppice biomass and an in-
tions were determined by a UV–vis spectrophotometer (Shimadzu UV- crease in activation energy, with a similar decrease in volatile yield but
1800) at 664 nm calibrated using 13 points over a concentration range a decrease in activation energy for a synthetic biomass mixture

98
L. Gao, J.L. Goldfarb Journal of Analytical and Applied Pyrolysis 137 (2019) 96–105

Table 2
Product yields (total solid, gas and liquid phases) for pyrolyzed samples as a function of starting sample mass (WS + FA) and on a per biomass yield (WS only) ± one
standard error.
Sample Wt % Pyrolysis Solid Yield (wt%) Gas Yield (wt%) Liquid Yielda (wt%) Solid Yield Gas Yield Liquid Yielda
Biomass Temperature (°C) (wt%) (wt%) (wt%)
Total Product Distribution [per mass sample Biomass Product Distribution [per mass
(WS + FA) basis] biomass (WS only) basis°]

Pyrolysis Temperature Experiments


WS_FA(20:1) 95 450 32.0 ± 0.1 19.0 ± 0.3 49.0 ± 0.1 28.4 20.1 51.5
WS_FA(20:1) 95 550 29.5 ± 0.2 36.2 ± 0.1 34.3 ± 0.2 25.8 38.1 36.1
WS_FA(20:1) 95 650 29.6 ± 0.2 41.7 ± 0.1 28.7 ± 0.2 25.9 43.9 30.2
WS_FA(20:1) 95 750 27.6 ± 0.2 43.4 ± 0.1 29.0 ± 0.2 23.8 45.7 30.5

WS:FA Ratio Experiments


WS 100 650 26.5 ± 0.2 36.8 ± 0.1 36.7 ± 0.1 26.5 36.8 36.6
WS_FA(100:1) 99 650 27.3 ± 0.2 39.8 ± 0.1 32.9 ± 0.2 26.5 40.2 33.3
WS_FA(50:1) 98 650 28.9 ± 0.2 40.5 ± 0.1 30.6 ± 0.1 22.7 44.0 33.3
WS_FA(20:1) 95 650 29.6 ± 0.2 41.7 ± 0.2 28.7 ± 0.2 25.9 43.9 30.2
WS_FA(10:1) 90 650 32.1 ± 0.2 33.1 ± 0.3 34.8 ± 0.2 24.6 36.8 38.6
FA 0 650 93.3 ± 0.1 5.0 ± 0.2 1.7 ± 0.2

°standard deviations equivalent to total product distribution.


a
by difference.

comprised of cellulose, lignin, organosolv and xylan from oat pellets heating. At 650 °C, considerably fewer phenols and substituted ben-
[33]. Other studies, such as micro- and nano-NiO catalyzed cellulose, zenes were detected. While the yield of phenol from biomass pyrolysis
xylan and lignin pyrolysis show minimal impact on reaction kinetics often increases at higher temperatures, in the presence of acidic cata-
[34]. While the lack of difference in reaction rate suggests that there is lysts phenols have been shown to alkylate [38]. Phenolic ethers are
minimal or no impact of incorporating the coal fly ash into the wheat often energetically favorable [39], though were not detected here in
straw, this is only one measurement on which to evaluate catalytic appreciable quantities. Given the higher pyrolysis temperatures, and
activity. corresponding increase in gas yields with increases in temperatures,
these compounds may well be subjected to thermal cracking, producing
CO and paraffins [40]. Others have shown that slow pyrolysis, even at
3.1. Impact of pyrolysis temperature on 5 wt% fly ash in wheat straw higher temperatures, tends to decrease the yields of acetic acid, phenol
(and methylated phenols), as well as other substituted aromatics [27],
The yields of pyrolysis gas, bio-oil and biochar at a series of po- as observed here.
tential pyrolysis temperatures (450 to 750 °C, at 100 °C increments) for In summation, while the total devolatilizations of the solid at 550
the WS/FA ratio of 20:1 were used to determine the temperature for and 650 °C were approximately equal, there was a significant shift in
bulk sample pyrolysis that maximized biofuel quality and quantity product composition, especially in the gas phase in terms of CO2 versus
while tempering energy input (by minimizing pyrolysis temperature). H2 and CH4 yields, suggesting that the higher temperature may be
As shown in Table 2, the total solid conversion increased from 450 to beneficial to bio-fuel production. While the yield of biochar at 650 °C
550 °C, then stayed roughly equivalent from 550 to 650 °C, and in- pyrolysis was only moderately higher than the yield at 750 °C (sug-
creased again when heated to 750 °C. However, if we look at the second gesting slightly lower conversion at 650 °C), it is not clear that in-
set of product distributions in Table 2, on a per mass of biomass basis, creasing the temperature would be beneficial to the overall conversion.
we find that while the biomass converted does indeed increase from In terms of the pyrolysis gas, the yields of C2H2, C2H4 and C2H6 were
550 to 650 °C, that this increases only the gas yield, and the liquid yield similar at 650 °C and 750 °C. Furfural, a promising platform chemical to
actually decreases from 550 to 650 °C. It is well documented that higher produce various chemicals via oxidation, hydrogenation and con-
pyrolysis temperatures generally result in lower solid char yields, and densation, was maximized at 650 °C, with minimal difference in 5-hy-
correspondingly higher gas yields as more of the solid is devolatilized droxymethylfurfural (another key intermediary) yield between 650 and
[35,36]. Yet, yield is only one part of the pyrolysis temperature deci- 750 °C. Therefore, to explore the impact of coal fly ash loading as a
sion; a survey of the condensable and non-condensable gases reveals the catalyst for pyrolysis of wheat straw, 650 °C was selected as a pyrolysis
impact of temperature on specific products. temperature to reduce energy input into the pyrolysis step while en-
From Table 3, we see that increasing pyrolysis temperature from abling a fairly complete devolatilization and conversion into desirable
450 °C to 750 °C (with a hold time of 60 min) dramatically increased the platform chemicals.
H2 yield by over 2800%. For C2H2, C2H4, C2H6, the yields increased
significantly from 450 to 550 °C, but then saw little increase in going
from 550 to 750 °C for each gas. As temperature increases, devolatili- 3.2. Impact of fly ash on wheat straw pyrolysis products at 650 °C
zation rate increases and thermal cracking leads to the formation of
more non-condensable gases [37]. As Table 2 shows, the inclusion of 1 wt% FA does not appreciably
As shown in Table 4, the majority of the condensable components change the overall conversion – the solid yield on a WS basis is 26.5%
identified were heterocyclic organics such as lactones (cyclic esters), for both WS and WS/FA(100:1). However, it does change the product
furans, phenols, and benzenediols, as well as alkenes, aldehydes, and distribution; the 1 wt% FA shifts the product distribution to the non-
carboxylic acids, commonly identified for pyrolysis of lignocellulose. In condensable gas phase. Interestingly, the specific gas component yields
Table 4 (chromatograms available in SI) we see identified compounds vary significantly with only 1 wt% FA in the biomass; as compared to
with their relative chromatogram area, normalized to pyrolyzed sample the pure WS, the H2 yield decreases by 15%, and C2H6 by 32%, whereas
mass and calibrated marker compounds with their concentration nor- the CO2, CH4 and C2H4 yields increase. This could be due to the lower
malized to pyrolyzed sample mass. As the pyrolysis temperature in- peak temperature of pyrolysis gas components evolved versus the raw
creased from 450 to 650 °C, the yield of furfural, a key biorefinery in- wheat straw, evolving more CO2 as is sometimes seen for low-tem-
termediate compound, increased, and then decreased upon further perature biomass pyrolysis [41]. It may also be attributed to the

99
L. Gao, J.L. Goldfarb

Table 3
Yield of pyrolysis gases ( ± 0.1 mggas/gWS) devolatilized at different temperatures and ash loadings normalized to wheat straw sample mass (WS only) and peak gas yield temperatures ( ± 1 °C; furnace K-type
thermocouple accuracy) as a function of ash loading.
Sample Wt % Pyrolysis H2 Rel CH4 Rel C2H2 Rel C2H4 Rel C2H6 Rel CO2 Rel
Biomass Temperature (°C) (mg/ change (mg/ change (mg/ change (mg/ change (mg/ change (mg/ change
gWS) (%) gWS) (%) gWS) (%) gWS) (%) gWS) (%) gWS) (%)

Pyrolysis Temperature Experiments - Pyrolysis Gas Yields


WS_FA(20:1) 95 450 0.2 8.1 0.8 9.5 0.6 54.0
WS_FA(20:1) 95 550 0.7 304% 10.9 35% 1.4 73% 12.8 36% 1.1 78% 152.2 182%
WS_FA(20:1) 95 650 2.1 1094% 15.7 94% 1.6 97% 15.7 66% 1.3 111% 167.6 210%
WS_FA(20:1) 95 750 5.2 2815% 17.8 119% 1.7 104% 14.5 54% 1.3 112% 193.9 259%

WS:FA Ratio Experiments - Pyrolysis Gas Yields


WS 100 650 2.4 13.5 1.7 12.6 1.6 143.9
WS_FA(100:1) 99 650 2.1 −15% 14.1 5% 1.5 −8% 12.7 0% 1.1 −32% 159.8 11%

100
WS_FA(50:1) 98 650 2.3 −4% 16.2 20% 1.7 2% 16.2 28% 1.4 −10% 173.1 20%
WS_FA(20:1) 95 650 2.1 −13% 15.7 16% 1.6 −1% 15.7 24% 1.3 −15% 167.6 16%
WS_FA(10:1) 90 650 2.1 −14% 14.4 7% 1.6 −5% 12.4 −2% 1.2 −22% 171.4 19%

WS:FA Ratio Experiments - Peak Pyrolysis H2 (°C) CH4 Peak 1 CH4 Peak 2 C2H2 (°C) C2H4 (°C) C2H6 (°C) CO2 (°C)
Gas Temperatures (°C) (°C)

WS 100 650 650 389 569 519 528 524 385


WS_FA(100:1) 99 650 650 383 565 519 519 520 380
WS_FA(50:1) 98 650 650 360 544 501 501 493 362
WS_FA(20:1) 95 650 633 331 517 465 474 464 329
WS_FA(10:1) 90 650 620 318 498 454 459 452 315
FA Pyrolysis at 650 °Ca H2 (mg/ H2 Peak CH4 (mg/gFA) CH4 Peak C2H2 (mg/ C2H2 Peak C2H4 (mg/ C2H4 Peak C2H6 (mg/ C2H6 Peak CO2 (mg/ CO2 Peak
gFA) (°C) (°C) gFA) (°C) gFA) (°C) gWS) (°C) gWS) (°C)
FA 0 650 1.00E-08 650 3.44E-07 206, 650 3.31E-08 – 2.43E-07 – 4.85E-08 – 104.0 202, 650
a
no peaks detected in analysis.
Journal of Analytical and Applied Pyrolysis 137 (2019) 96–105
L. Gao, J.L. Goldfarb Journal of Analytical and Applied Pyrolysis 137 (2019) 96–105

Table 4
Condensable components resulting from pyrolysis of WS/FA (20:1) at a series of different temperatures as identified via GCMS; Calibrated compounds given in
mgcompound per gWS ( ± 0.0005 mg/g).
Retention time (min) Compound 450 °C 550 °C 650 °C 750 °C
Normalized Concentration (mg/gWS)

11.6 Furfural [2-furancarboxaldehyde] 8.84E-02 2.40E-01 2.48E-01 2.01E-01


13.69 2-furanmethanol 7.91E-02 1.68E-01 8.60E-02 9.54E-02
15.79 Styrene [Ethenylbenzene] 8.02E-03 ND ND ND
17.98 2(5 H)-furanone 1.61E-02 4.03E-02 ND 8.71E-03
18.23 1,2-cyclopentanedione 1.15E-02 1.36E-02 ND 4.26E-03
21.03 5-hydroxymethyl furfural 8.21E-03 2.21E-02 5.99E-03 9.08E-03
22.43 Phenol 1.31E-02 3.28E-03 2.01E-03 3.22E-03
24.73 3-methyl-1,2-cyclopentanedione 3.48E-03 4.32E-03 ND 2.48E-03
27.73 p-cresol [4-methylphenol] 5.10E-03 1.53E-03 ND 1.16E-03
34.03 Catechol [1,2-dihydroxybenzene] 1.31E-03 ND ND 2.44E-03
40.93 Syringol [2,6-dimethoxyphenol] 1.62E-03 3.00E-03 ND 2.77E-03

Normalized Chromatogram Area (TIC/g)


7.94 Butanedial 2.01E+05 9.95E+04 4.10E+04 8.43E+04
8.35 (2 H)-furan-2-one 1.78E+04 3.67E+04 ND 1.62E+04
8.58 Butanoic acid ND 2.68E+04 1.99E+05 3.01E+04
8.77 Acetic acid ND 2.44E+05 1.10E+05 2.02E+05
9.25 2-hexene ND 3.30E+04 ND 2.21E+04
11.46 2-cyclopenten-1-one ND 4.80E+05 8.89E+05 8.22E+05
14.1 Ethylbenzene 1.00E+05 ND ND ND
17.52 Ethanone,1-(2-furanyl) 5.51E+04 1.42E+05 1.35E+05 1.47E+05
17.76 Butyrolactone 3.50E+04 1.15E+05 2.47E+04 1.97E+04
28.21 2-methoxy-phenol 9.09E+04 2.76E+04 ND 1.75E+04
28.41 Pentanal 1.52E+05 1.04E+05 4.83E+04 1.72E+05
34.27 1,4:3,6-dianhydro-,alpha,-d-glucopyranose 1.03E+04 2.64E+04 ND 2.79E+04

*
ND = not detected.

Fig. 2. Mass spectra of H2 and CH4 evolved during 650 °C pyrolysis of wheat straw and wheat straw + fly ash blends (spectra of all monitored gases available in
Supplemental Information).

mineral matter of the biomass and/or coal producing CO2 through devolatilization of the FA alone, but must be due to an interaction
carbonate decomposition and/or interactions between the mineral between the FA and WS. As shown in Table 3 and Fig. 2 (additional
matter and biomass carbon, which might explain the increase in CH4 plots available in SI), the peak temperature at which various com-
and C2H4 yields [42]. This is supported by the pyrolysis product dis- pounds were produced decreased upon incorporation of the 1 wt% FA.
tribution of the coal fly ash. When FA is pyrolyzed at 650 °C, the only These peak evolution temperatures decreased considerably as the
compound to be evolved in any appreciable quantity is CO2 (Table 3). amount of FA included in the biomass increased, as shown in Table 3.
Thus, the higher yields of CH4 and C2H4 are not attributable to the For many gases, including CH4, the gases were evolved at faster rates as

101
L. Gao, J.L. Goldfarb Journal of Analytical and Applied Pyrolysis 137 (2019) 96–105

the FA increased, suggesting that the coal fly ash may catalyze the fly ash in the blend increased. For example, two compounds containing
devolatilization of biomass samples. eC]O, 2(5 H)-Furanone and 1,2-Cyclopentanedione, were removed
For the WS/FA (10:1) sample, the inclusion of coal fly ash into the completely when the coal fly ash to wheat straw ratio is (1:20) or
wheat straw has a significant impact on methane devolatilized, in- higher; interestingly the former compound peaked at the (100:1) ratio,
creasing the total yield (versus pure WS) by 7%, but decreasing the while the latter peaked for the pure WS, and decreased by 50% for
peak evolution temperature by almost 70 °C. The peak temperatures of (100:1) blend. The concentrations of acidic compounds displayed
C2H4 and C2H6 release also decrease by approximately 70 °C. somewhat divergent behavior. For propanoic acid, the yield increased
Interestingly, the peak temperature of each evolved pyrolysis gas with FA up to (20:1), then disappeared for the (10:1) blend; likewise,
component shifts in a to similar lower temperatures as a function of FA for acetic acid the concentration peaked at (100:1) and was not de-
loading, with the exception of hydrogen. For example, every gas com- tected for (10:1). Conversely, butanoic acid saw a fairly similar yield for
ponent devolatilized at approximately 50, 20, and 5 °C lower tem- WS, (100:1), (50:1) and (20:1) blends, but was not detected in the
peratures than WS for WS/FA (20:1), (50:1), and (100:1) blends, re- (10:1) blend. The catalytic pyrolysis of biomass often proceeds via the
spectively. As compared to the dramatic increase in gas yields seen for breaking of CeCO(OH) and CeOH bonds in decarboxylation, dec-
different pyrolysis temperatures, the amounts of gaseous fuels were arbonylation and dehydration reactions [8]. It appears that the fly ash
only modestly impacted by FA loading. Overall, the hydrogen evolved may deoxygenate condensable liquids in terms of lowering the yield of
decreased for each sample including FA, from a low of 4% for the WS/ cyclic oxygenated groups even at low loadings, but requires more sig-
FA (50:1) blend to 13–15 % for the other blends. Likewise, C2H6 de- nificant quantities to reduce the acidic components. While we do not
creased for all samples. However, methane increased slightly upon in- have the ability to monitor CO production (as the reaction was run in
corporation of fly ash into the biomass, up to 20% for the WS/FA (50:1) N2, and both show the same MS signal), the fluctuating levels of CO2
blend. Longer reaction times and higher temperatures tend to favor devolatilized mimic the overall results here; while there is no specific
recombination of smaller radicals into large gas phase molecules, “trend” for CO2, the additional of coal fly ash increases its yield
whereas lower temperatures suppress thermal cracking. The behavior (overall) while decreasing (on average) the oxygenated condensable
observed might be an artifact of gases evolving at lower temperatures yield.
and shorter residence times [43]. While the fly ash appears to have a Furfural (2-furancarboxaldehyde) is a valuable platform chemical
small catalytic impact on the overall gaseous products produced – for the production of pharmaceuticals, resins, fungicides and other
especially at lower loadings (1 wt% data) – at higher loadings it can specialty chemicals [46]. Furfural is the most common and dominant
significantly lower the peak evolution temperatures and, in most cases, intermediate compound during biomass pyrolysis, but its yield is
increase the evolution rate at which the gas is devolatilized. Given the usually lower than 1 wt% [47]. Previously it was shown that the fur-
lack of impact of the FA on the DTG curves this was somewhat sur- fural yield could be promoted by acid catalyzed pyrolysis processes
prising. One might suppose that the peak mass loss would correspond to using catalysts such as ZnCl2, MgCl2, CaCl2, AlMCM-41, (NH)2HPO4,
gaseous component evolution. However, this is not always the case; as Fe2(SO4)3, and Al/SBA-15 [48–58]. From Fig. 3, we see that as the
noted throughout the biomass pyrolysis literature, some “directly de- amount of coal fly ash increases from 1 wt% to 5 wt%, the furfural yield
volatilized” compounds (e.g. those compounds that can be direct pro- increases by 66 then 94% (almost doubling for 5 wt% FA), then falls
ducts of thermal degradation such as H2O and CO2) do correspond to back to only a 40% increase over raw wheat straw for the 10 wt% fly
peak DTG curves, but for other gaseous components, they may occur ash blend. Results obtained in the present research suggest that the
before/after peak mass loss [44,45] as is the case here. As such, we turn weight ratio of coal fly ash to wheat straw of (20:1) is effective in both
to an analysis of the condensable bio-oil via GC–MS to shed light on this increasing the furfural yield while decreasing some oxygenated com-
observation and to further gauge the efficacy of coal fly ash to act as a ponents. However, this is a balancing act of desirable components, as
catalyst for the pyrolysis of biomass to biofuels. the (100:1) blend significantly increased the 5-hydroxymethyl furfural
Fig. 3 compares the bio-oil components identified via GCeMS yield over the raw wheat straw, but at higher FA loadings the 5-HMF
analysis of the condensable bio-fuel trapped in methylene chloride yield returned to that seen without the fly ash.
(DCM) during pyrolysis of the WS and WS + FA blends at 650 °C Previous results showed that activated fly ash and materials ex-
(chromatograms and data provided in SI). Again, the primary con- tracted from fly ash are promising catalysts for various biomass reac-
densable compounds identified were furfurals, carboxylic acids and tions, especially for extraction of minerals and for the production of
esters. As shown in Fig. 3, increasing the coal fly ash positively impacts zeolites [59]. Here we find coal fly ash itself has a catalytic effect on
the yield of compounds with oxygen containing functional groups. That wheat straw pyrolysis. Evidence for this includes: (1) increased devo-
is, compounds with eOH, eCOOH, eC]O, eCeOeC groups were latilization of biomass (lower char yield) at same temperature/re-
overall removed – to varying degrees and amounts – as the amount of sidence times favoring gaseous (noncondensable) products as catalyst

Fig. 3. Yield of condensable components as a function of fly ash loading for pyrolysis at 650 °C.

102
L. Gao, J.L. Goldfarb Journal of Analytical and Applied Pyrolysis 137 (2019) 96–105

Table 5
Characterization and application of heterogeneous WS + FA biochars prepared at 650 °C for adsorption of methylene blue from water.
Biochar Volatile matter (wt%) Fixed carbon (wt%) BET surface area (m2/g) Pore volume (cm3/g)

WS 22.73 ± 1.27 62.24 ± 0.65 5.9628 ± 0.13 0.0121 ± 0.0001


WS/FA(100:1) 17.65 ± 0.59 64.48 ± 1.67 10.397 ± 0.057 0.0189 ± 0.0001
WS/FA(50:1) 16.50 ± 1.70 65.63 ± 1.94 14.457 ± 0.132 0.0212 ± 0.0001
WS/FA(20:1) 23.38 ± 2.14 56.03 ± 2.82 21.848 ± 0.36 0.0256 ± 0.0002
WS/FA(10:1) 12.61 ± 0.68 63.02 ± 1.05 5.6655 ± 0.17 0.0119 ± 0.0001
FA 0.55 ± 0.02 18.18 ± 1.27 36.173 ± 1.33 0.0302 ± 0.0001

Langmuir isotherm Freundlich isotherm

Biochar qm(mg/g) KL(L/mg) R2 Kf [mg/g(L/mg)1/n] n R2

WS 32.4 ± 1.08 2.87E-02 ± 2.90E-03 0.985 5.60 ± 0.01 3.36 ± 0.03 0.991
WS/FA(100:1) 32.6 ± 0.19 2.10E-02 ± 1.27E-03 0.993 3.10 ± 0.03 2.46 ± 0.07 0.973
WS/FA(50:1) 44.2 ± 0.19 8.48E-03 ± 1.22E-03 0.991 1.94 ± 0.01 2.01 ± 0.02 0.999
WS/FA(20:1) 54.8 ± 0.34 5.35E-03 ± 1.32E-03 0.968 1.30 ± 0.05 1.73 ± 0.11 0.976
WS/FA(10:1) 64.7 ± 0.24 1.19E-02 ± 1.84E-03 0.934 3.01 ± 0.02 1.94 ± 0.04 0.992
FA 41.6 ± 0.05 2.99E-02 ± 4.84E-04 0.998 6.30 ± 0.02 2.97 ± 0.01 0.997

Pseudo-first-order kinetics Pseudo-second-order kinetics

Biochar K1(1/min) R2 K2(g/mg min) R2

WS 5.69E-03 ± 1.05E-04 0.961 1.63E-03 ± 2.02E-04 0.993


WS/FA(100:1) 5.83E-03 ± 2.58E-04 0.993 3.01E-03 ± 1.27E-04 0.999
WS/FA(50:1) 9.13E-03 ± 6.35E-04 0.963 3.15E-03 ± 2.23E-04 0.997
WS/FA(20:1) 4.39E-03 ± 3.52E-04 0.951 3.68E-03 ± 2.46E-04 0.996
WS/FA(10:1) 5.36E-03 ± 4.29E-04 0.946 3.73E-03 ± 2.31E-04 0.996
FA 4.96E-03 ± 5.82E-04 0.890 1.08E-02 ± 3.52E-04 1.000

loading increases up to 5 wt% with shift to condensable products at Where x is the mass fraction of the component in the blend (predicted
10 wt% fly ash; (2) enhanced production of some pyrolysis gases, in- pore volume calculated analogously). In fact, the measured surface area
cluding CO2, CH4, and C2H4, with significant lowering of peak gas re- and pore volume were almost twice these values. This may due to the
lease temperatures (up to 70 °C) as the fraction of fly ash increases; and coal fly ash promoting devolatilization and removal of the pyrolysis
(3) higher yields of furfurals and lower of oxygenated compounds in the products from the solid, opening pores in the material. The surface area
condensable bio-oil fraction. These points suggest that the coal fly ash and pore volume of WS/FA(10:1) were almost the same as that of the
could act as a catalyst in biomass pyrolysis, especially for dehydration wheat straw.
and deoxygenation reactions. Beyond the catalytic benefits of in- Equilibrium adsorption studies on the ability of biochars to remove
corporating coal fly ash into biomass during pyrolysis, we explored the methylene blue (MB) from water show maximum adsorption capacities
potential waste-to-byproduct conversion of the heterogeneous biochars from 33 to 65 mgMB/gBiochar as determined via Langmuir isotherms
produced for their application to water treatment. (Table 5). The adsorption intensity of the Freundlich isotherm, n, values
are all greater than 1. This suggests that it is thermodynamically fa-
vorable to use these heterogeneous sorbents to remove methylene blue
3.3. Characterization and application of heterogeneous biochars
from aqueous solutions [60]. Given the reasonable fits of each model to
the data (with 15 data points R2 > 0.95 for every sample/model) it is
Biomass samples mixed with coal fly ash and pyrolyzed at 650 °C
difficult to assign specific isotherm behavior based on these models.
yielded higher surface areas than the biochar alone, with the exception
Langmuir is often thought to describe monolayer adsorption, whereby
of the (10:1) sample (Table 5). Interestingly, this sample had a similar
an equilibrium saturation point exists when the surface is saturated by
surface area and porosity to the biochar, despite the finding that the
chemically adsorbed species. The Freundlich isotherm is not con-
pyrolyzed fly ash has the highest surface area of all the samples. The
strained by this monolayer behavior, but rather describes reversible
porous voids that formed during devolatilization could be blocked by
physisorption [61]. This ambiguity in model selection extends to the
the coal fly ash or may have collapsed (plausible given high volume of
adsorption kinetic studies.
gas released and small amount of volatile matter present), both of
The adsorption of MB onto all biochars appears to “follow” pseudo-
which would lead to lower surface area and pore volume [17].
second-order kinetics as indicated by correlation coefficients
When the ratio of fly ash to wheat straw is below 1:20, both the
(R2 > 0.99) for all samples, a criterion often applied to deciding model
surface area and pore volume increase as the proportion of coal fly ash
“fit.” (Table 5, Fig. S4 of the Supporting Information). However, with
in the blend increases; for the (20:1) sample, the increase is approxi-
an R2 > 0.94 for pseudo-first-order models and 15 data points, the as-
mately four times higher than the biomass alone. The enhancement in
signment of an adsorption mechanism according to the kinetics model
surface area and pore volume as a function of coal fly ash addition is not
using R2 values is not statistically appropriate. The inconclusive nature
an additive function of the coal fly ash’s presence. For example, if the
of the rate law is common throughout the sorption literature for coal fly
coal fly ash had no effect on the biochar during pyrolysis but simply was
ash and fly ash + biomass sorbents [62]. We suspect that this is because
“present” in terms of mass fraction, we might expect the resulting WS/
the adsorption is both dependent on bulk MB concentration and capa-
FA(100:1) surface area and pore volume to be 6.265 m2/g and 0.0123
city (first-order) and on chemisorption involving valency forces be-
cm3/g if we assume that the contribution is a weighted function of the
tween sorbent and sorbate (second-order). Given the heterogeneity of
blends’ constituents as:
the biomass and, more importantly, the coal fly ash (as described in
SApredict= wWSSAWS + washSAash (5) Table 1), the potential for sorption to be governed by both diffusion and

103
L. Gao, J.L. Goldfarb Journal of Analytical and Applied Pyrolysis 137 (2019) 96–105

surface complexation – and, as noted above, by both physical and suggesting synergistic, catalytic behavior of the coal fly ash in pro-
chemical forces – is not surprising. Such concerted adsorption forces moting devolatilization. Overall, we find that coal fly ash may be a
have been observed throughout the literature [62]. potential inexpensive catalyst for in situ upgrading of biomass pyrolysis,
Prior research suggests that high surface area and pore volume are producing heterogeneous sorbents with enhanced adsorption capacities
linked to greater adsorption of pollutants onto solid adsorbents [20,63], for organic pollutants.
however these are not directly correlated, as recent research shows
[64]. The WS/FA(10:1) sample, with moderately lower surface area and Acknowledgements
pore volume than other biochars used in this study, shows moderately
higher adsorption capacities than the other samples, as seen in Table 5 L. Gao acknowledges the support of the Fundamental Research
(and Fig. S5 of Supplemental Information). The WS/FA (20:1) and Funds for the Central Universities (No. 2017BSCXA15), the
(10:1) samples have a faster rate of adsorption compared with the other Postgraduate Research & Practice Innovation Program of Jiangsu
biochars, especially the raw WS. The WS/FA(10:1) sample, with a Province (No. KYCX17_1515) and the China Scholarship Council. The
surface area and pore volume of 5.66 m2/g and 0.0119 cm3/g, had the authors appreciate the assistant of Quantachrome with surface area
highest adsorption capacity and fastest adsorption kinetic rate for the analysis of the fly ash. The authors sincerely thank the anonymous
methylene blue, of 64.715 mg/g and 3.73E-3 g/(mg·min), as compared Reviewers who significantly improved this manuscript.
to the almost same surface area and pore volume of pure WS with an
adsorption capacity and rate constant of 32.398 mg/g and 1.63E-3 g/ Appendix A. Supplementary data
(mg·min). Sample WS/FA(20:1), with a surface area and pore volume of
21.848 m2/g and 0.0256 cm3/g, had a moderate lower capacity and Supplementary material related to this article can be found, in the
rate (54.839 mg/g and 3.68E-3 g/(mg·min) than WS/FA(10:1). online version, at doi:https://doi.org/10.1016/j.jaap.2018.11.014.
The addition of the coal fly ash to the biomass also shows synergistic
effects (rather than additive) in terms of adsorption capacities. Just as References
the relationship between surface area and fraction of coal fly ash pre-
sent in the mixture was not additive (per Eq. (5)), neither is the ad- [1] M.F. Othman, A. Adam, G. Najafi, R. Mamat, Green fuel as alternative fuel for diesel
sorption capacity. The pyrolyzed fly ash had a higher adsorption ca- engine: a review, Renew. Sustain. Energy Rev. (December) (2017) 694–709
Pergamon.
pacity than the biochar pyrolyzed at the same conditions, and for the [2] M. Rydén, A. Lyngfelt, O. Langørgen, Y. Larring, A. Brink, S. Teir, H. Havåg,
(100:1) sample the fly ash had no measurable impact on the adsorption P. Karmhagen, Negative CO2 emissions with chemical-looping combustion of bio-
capacity (or rate). However, again we see that as the proportion of fly mass - a nordic energy research flagship project, Energy Procedia vol. 114, Elsevier,
2017, pp. 6074–6082.
ash increases, the adsorption capacity increases – but this time con- [3] L.Q. Ji, C. Zhang, J.Q. Fang, Economic analysis of converting of waste agricultural
siderably higher than might be predicted by an additive relationship. biomass into liquid fuel: a case study on a biofuel plant in China, Renew. Sustain.
Using Eq. (5), we would expect the adsorption capacity of all the fly ash Energy Rev. (April) (2017) 224–229 Pergamon.
[4] A. Oasmaa, S. Czernik, Fuel oil quality of biomass pyrolysis oil - state of the art for
containing mixtures to be around 32–33 mg/g. However, the adsorption
the end users, Energy Fuels 13 (4) (1999) 914–921.
capacities were clearly much higher. This suggests, just as observed for [5] K. Iisa, D.J. Robichaud, M.J. Watson, J. ten Dam, A. Dutta, C. Mukarakate, S. Kim,
the surface area enhancements, bio-oil quality enhancements, and M.R. Nimlos, R.M. Baldwin, Improving biomass pyrolysis economics by integrating
vapor and liquid phase upgrading, Green Chem. 20 (3) (2018) 567–582.
pyrolysis gas yield increases that the fly ash acts as a catalytic material
[6] L. Wang, F.-S. Xiao, Nanoporous catalysts for biomass conversion, Green Chem. 17
to increase devolatilization and improve the porous nature of the re- (1) (2015) 24–39.
sulting heterogeneous adsorbents. [7] S. Xiu, A. Shahbazi, Bio-oil production and upgrading research: a review,
Renewable and Sustainable Energy Reviews, Elsevier, 2012, pp. 4406–4414.
[8] Y. Lin, G. Huber, The critical role of heterogeneous catalysis in lignocellulosic
4. Conclusions biomass conversion, Energy Environ. Sci. (2009).
[9] J. Jae, R. Coolman, T.J. Mountziaris, G.W. Huber, Catalytic fast pyrolysis of lig-
Solid waste disposal – be it coal fly ash or residual biomasses – nocellulosic biomass in a process development unit with continual catalyst addition
and removal, Chem. Eng. Sci. 108 (2014) 33–46.
strains the environment in many ways, and may well be mitigated [10] J. Xue, G. Dou, E. Ziade, J.L. Goldfarb, Integrating sustainable biofuel and silver
through the conversion of waste to renewable energy and sustainable nanomaterial production for in situ upgrading of cellulosic biomass pyrolysis,
materials for water treatment. The present study demonstrates the Energy Convers. Manage. 142 (2017).
[11] J. Gupta, K. Papadikis, I.V. Kozhevnikov, E.Y. Konysheva, Exploring the potential of
possibility of using fly ash from the combustion of a bituminous coal to red mud and Beechwood Co-Processing for the upgrading of fast pyrolysis vapours,
in situ catalyze the pyrolysis of wheat straw, an abundant agricultural J. Anal. Appl. Pyrolysis 128 (2017) 35–43.
biomass. The inclusion of coal fly ash did not significantly impact the [12] R. French, S. Czernik, Catalytic pyrolysis of biomass for biofuels production, Fuel
Process. Technol. 91 (1) (2010) 25–32.
overall reaction rate, but it did lower the temperature at which many [13] R.S.C. Advances, S. Dutta, Catalytic materials that improve selectivity of biomass
noncondensable gaseous species were released (potentially lowering conversions, RSC Adv. 2 (33) (2012) 12575.
the energy required to pyrolyze more of the biomass), and resulted in a [14] P. Rutkowski, Pyrolytic behavior of cellulose in presence of montmorillonite K10 as
catalyst, J. Anal. Appl. Pyrolysis 98 (2012) 115–122.
condensable liquid with fewer detectable oxygenated components and
[15] Tkun Huang, R. Wang, L. Shi, Xxia Lu, Montmorillonite K-10: an efficient and
an increase in furan, a key biorefinery intermediary. As the weight reusable catalyst for the synthesis of quinoxaline derivatives in water, Catal.
fraction of fly ash in the biomass increased from 1 to 10 wt%, the Commun. 9 (6) (2008) 1143–1147.
overall conversion of solid biomass increased; 1–5 wt% fly ash favored [16] N. Kaur, D. Kishore, Montmorillonite: an efficient, heterogeneous and green catalyst
for organic synthesis, J. Chem. Pharm. Res. (2012) 991–1015.
conversion to the gas phase (especially CO2, CH4 and C2H4), whereas [17] G. Dou, J. Goldfarb, In situ upgrading of pyrolysis biofuels by bentonite clay with
the 10 wt% mixture produced more condensable species. For a more simultaneous production of heterogeneous adsorbents for water treatment, Fuel 195
complete waste-to-byproduct transformation, we probed the ability of (2017) 273–283.
[18] B.G. Kutchko, A.G. Kim, Fly ash characterization by SEM-EDS, Fuel 85 (17–18)
the resulting heterogeneous wheat straw-fly ash biochars to removed (2006) 2537–2544.
methylene blue from water as a simulated organic pollutant. Like many [19] S. Wang, Application of solid ash based catalysts in heterogeneous catalysis.
heterogeneous biomass-mineral systems, there is ambiguity in de- Environmental science and technology, Am. Chem. Soc. (October) (2008)
7055–7063.
termining whether or not the adsorption is physically or chemically [20] S. Li, L. Gao, H. Wen, G. Li, Y. Wang, Modification and application of coking coal by
motivated; likely it is a combination of mechanisms given the hetero- alkali pretreatment in wastewater adsorption, Sep. Sci. Technol. 52 (16) (2017)
geneity of the mineral matter comprising both the biomass and coal fly 2532–2539.
[21] J. Han, H. Kim, The reduction and control technology of tar during biomass
ash. However, the incorporation of the fly ash significantly improves Gasification/Pyrolysis: an overview, Renew. Sustain. Energy Rev. (February)
the surface area and adsorption capacity of the heterogeneous biochars, (2008) 397–416.
above what might be predicted from an additive relationship, [22] Y.C. Rotliwala, P.A. Parikh, Thermal Co-processing of high density polyethylene

104
L. Gao, J.L. Goldfarb Journal of Analytical and Applied Pyrolysis 137 (2019) 96–105

with coal, fly ashes, and biomass: characterization of liquid products, Energy [44] J. Xue, G. Dou, E. Ziade, J.L. Goldfarb, Integrating sustainable biofuel and silver
Sources, Part A Recover. Util. Environ. Eff. 34 (11) (2012) 1055–1066. nanomaterial production for in situ upgrading of cellulosic biomass pyrolysis,
[23] P. Gaurh, H. Pramanik, Production of Benzene/Toluene/Ethyl Benzene/Xylene Energy Convers. Manage. 142 (2017) 143–152.
(BTEX) via multiphase catalytic pyrolysis of hazardous waste polyethylene using [45] G. Dou, J.L. Goldfarb, In situ upgrading of pyrolysis biofuels by bentonite clay with
low cost fly ash synthesized natural catalyst, Waste Manag. 77 (2018) 114–130. simultaneous production of heterogeneous adsorbents for water treatment, Fuel 195
[24] L. Zhang, K. Chen, L. He, L. Peng, Reinforcement of the bio-gas conversion from (2017) 273–283.
pyrolysis of wheat straw by hot caustic pre-extraction, Biotechnol. Biofuels 11 (1) [46] Q. Lu, C.Q. Dong, X.M. Zhang, H.Y. Tian, Y.P. Yang, X.F. Zhu, Selective fast pyr-
(2018) 72. olysis of biomass impregnated with ZnCl2 to produce furfural: analytical Py-GC/MS
[25] P.A. Jensen, B. Sander, K. Dam-Johansen, Pretreatment of straw for power pro- study, J. Anal. Appl. Pyrolysis 90 (2) (2011) 204–212.
duction by pyrolysis and char wash, Biomass Bioenergy 20 (6) (2001) 431–446. [47] C. Di Blasi, C. Branca, A. Galgano, Biomass screening for the production of furfural
[26] B. Chen, X. Han, M. Mu, X. Jiang, Studies of the Co-pyrolysis of oil shale and wheat via thermal decomposition, Ind. Eng. Chem. Res. 49 (6) (2010) 2658–2671.
straw, Energy Fuels 31 (7) (2017) 6941–6950. [48] F. Shafizadeh, Y.Z. Lai, Thermal degradation of 1,6-Anhydro-.Beta.-D-
[27] C.J. Mulligan, L. Strezov, V. Strezov, Thermal decomposition of wheat straw and Glucopyranose, J. Org. Chem. 37 (2) (1972) 278–284.
mallee residue under pyrolysis conditions, Energy Fuels 24 (2010) 46–52. [49] A. Ramiro, J.F. Gonza, Catalyzed pyrolysis of grape and olive bagasse. Influence of
[28] J. Xue, T. Chellappa, S. Ceylan, J.L. Goldfarb, Enhancing biomass + coal Co-firing catalyst type and chemical treatment, Ind. Eng. Chem. Res. 36 (10) (1997)
scenarios via biomass torrefaction and carbonization: case study of avocado pit 4176–4183.
biomass and illinois No. 6 coal, Renew. Energy 122 (2018) 152–162. [50] P.R. Patwardhan, J.A. Satrio, R.C. Brown, B.H. Shanks, Influence of inorganic salts
[29] A.S. Patnaik, J.L. Goldfarb, Continuous activation energy representation of the ar- on the primary pyrolysis products of cellulose, Bioresour. Technol. 101 (12) (2010)
rhenius equation for the pyrolysis of cellulosic materials: feed corn stover and cocoa 4646–4655.
shell biomass, Cellul. Chem. Technol. 50 (2) (2016). [51] E. Me, M. Sto, J. Adam, M. Blazso, A. Bouzga, J.E. Hustad, M. Grønli, G. Øye,
[30] R. Evans, T. Milne, Molecular characterization of the pyrolysis of biomass, Energy Pyrolysis of biomass in the presence of Al-MCM-41 type catalysts, Fuel 84 (12–13)
Fuels 1 (2) (1987) 123–137. (2005) 1494–1502.
[31] Y.F. Huang, W.H. Kuan, P.T. Chiueh, S.L. Lo, Pyrolysis of biomass by thermal [52] C. Blasi, Di, C. Branca, A. Galgano, I. Chimica, V. Uni, F. Ii, P.V. Tecchio, Effects of
analysis – mass spectrometry (TA – MS), Bioresour. Technol. 102 (2011) diammonium phosphate on the yields and composition of products from wood
3527–3534. pyrolysis, Ind. Eng. Chem. Res. 46 (2) (2007) 430–438.
[32] S. Vyazovkin, K. Chrissafis, M.L. Di Lorenzo, N. Koga, M. Pijolat, B. Roduit, [53] Mqiang Chen, J. Wang, Mxu Zhang, Mgong Chen, Xfeng Zhu, Ffei Min, Zcheng Tan,
N. Sbirrazzuoli, J.J. Suñol, ICTAC kinetics committee recommendations for col- Catalytic effects of eight inorganic additives on pyrolysis of pine wood sawdust by
lecting experimental thermal analysis data for kinetic computations, Thermochim. microwave heating, J. Anal. Appl. Pyrolysis 82 (1) (2008) 145–150.
Acta 590 (2014) 1–23. [54] C. Blasi, Di, C. Branca, A. Galgano, Products and global weight loss rates of wood
[33] D.J. Nowakowski, J.M. Jones, R.M.D. Brydson, A.B. Ross, Potassium catalysis in the decomposition catalyzed by zinc chloride, Energy Fuels 22 (1) (2008) 663–670.
pyrolysis behaviour of short rotation willow coppice, Fuel 86 (15) (2007) [55] N. Shimada, H. Kawamoto, S. Saka, Different action of Alkali/Alkaline earth metal
2389–2402. chlorides on cellulose pyrolysis, J. Anal. Appl. Pyrolysis 81 (1) (2008) 80–87.
[34] J. Li, R. Yan, B. Xiao, D. Liang, D. Lee, Preparation of Nano-NiO particles and [56] L. Qiang, L. Wen-zhi, Z. Dong, Z. Xi-feng, Analytical pyrolysis-gas Chromatography/
evaluation of their catalytic activity in pyrolyzing biomass components†, Energy Mass spectrometry (Py-GC/MS) of sawdust with Al/SBA-15 catalysts, J. Anal. Appl.
Fuels 333 (Part 2) (2007) 16–23. Pyrolysis 84 (2) (2009) 131–138.
[35] E.C. Efika, J.A. Onwudili, P.T. Williams, Products from the high temperature pyr- [57] Q. Lu, W.M. Xiong, W.Z. Li, Q.X. Guo, X.F. Zhu, Catalytic pyrolysis of cellulose with
olysis of RDF at slow and rapid heating rates, J. Anal. Appl. Pyrolysis (2015) 14–22. sulfated metal oxides: a promising method for obtaining high yield of light furan
[36] D. Chen, L. Yin, H. Wang, P. He, Reprint of: pyrolysis technologies for municipal compounds, Bioresour. Technol. 100 (20) (2009) 4871–4876.
solid waste: a review, Waste Management, Elsevier Ltd, 2015, pp. 116–136. [58] Y. Wan, P. Chen, B. Zhang, C. Yang, Y. Liu, X. Lin, R. Ruan, Microwave-assisted
[37] A. Sharma, V. Pareek, D. Zhang, Biomass pyrolysis - a review of modelling, process pyrolysis of biomass: catalysts to improve product selectivity, J. Anal. Appl.
parameters and catalytic studies, Renew. Sustain. Energy Rev. (2015) 1081–1096 Pyrolysis 86 (1) (2009) 161–167.
Elsevier. [59] C. Belviso, State-of-the-Art applications of fly ash from coal and biomass: a focus on
[38] C.K. Rasulov, K.D. Ibragimov, E.A. Medzhidov, S.Z. Alieva, A.S. Askerova, Reaction zeolite synthesis processes and issues, Prog. Energy Combust. Sci. (March) (2018)
between phenol and 130-190°C fraction of liquid pyrolysis products in the presence 109–135 Pergamon.
of catalyst KU-23 in a continuous plant, Russ. J. Appl. Chem. 86 (1) (2013) 36–41. [60] J.L. Goldfarb, L. Buessing, E. Gunn, M. Lever, A. Billias, E. Casoliba, A. Schievano,
[39] Q. Ma, D. Chakraborty, F. Faglioni, R.P. Muller, W.A. Goddard, T. Harris, F. Adani, Novel integrated biorefinery for olive mill waste management: utilization
C. Campbell, Y. Tang, Alkylation of phenol: a mechanistic view, J. Phys. Chem. A of secondary waste for water treatment, ACS Sustain. Chem. Eng. 5 (1) (2017)
(2006) 2246–2252. 876–884.
[40] K. Brezinsky, M. Pecullan, I. Glassman, Pyrolysis and oxidation of phenol, J. Phys. [61] K.Y. Foo, B.H. Hameed, Insights into the modeling of adsorption isotherm systems,
Chem. 102 (1998) 8614–8619. Chem. Eng. J. 156 (1) (2010) 2–10.
[41] J. Xue, S. Ceylan, J.L. Goldfarb, Synergism among biomass building blocks? Evolved [62] Y. Ho, G. McKay, Pseudo-second order model for sorption processes, Process
gas and kinetics analysis of starch and cellulose Co-pyrolysis, Thermochim. Acta Biochem. 34 (5) (1999) 451–465.
618 (2015) 36–47. [63] L. Gao, H. Wen, Q. Tian, Y. Wang, G. Li, Influence of surface modification by sul-
[42] N. Ellis, M.S. Masnadi, D.G. Roberts, M.A. Kochanek, A.Y. Ilyushechkin, Mineral furic acid on coking coal’s adsorption of coking wastewater, Water Sci. Technol. 76
matter interactions during Co-pyrolysis of coal and biomass and their impact on (3) (2017) 555–566.
intrinsic char Co-gasification reactivity, Chem. Eng. J. 279 (2015) 402–408. [64] M. Berger, J. Ford, J.L. Goldfarb, Modeling aqueous contaminant removal due to
[43] F. Paradela, F. Pinto, A.M. Ramos, I. Gulyurtlu, I. Cabrita, Study of the slow batch combined hydrolysis and adsorption: oxytetracycline in the presence of biomass-
pyrolysis of mixtures of plastics, tyres and forestry biomass wastes, J. Anal. Appl. based activated carbons, Sep. Sci. Technol. (2018) 1–17.
Pyrolysis 85 (1–2) (2009) 392–398.

105

Anda mungkin juga menyukai