Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

View Article Online

View Journal

Green Chemistry

Accepted Manuscript
This article can be cited before page numbers have been issued, to do this please use: S. Du, J. Valla and G. M. Bollas, Green
Chem., 2013, DOI: 10.1039/C3GC41581C.

This is an Accepted Manuscript, which has been through the RSC Publishing peer
review process and has been accepted for publication.

Cutting-edge research for a greener sustainable future


Accepted Manuscripts are published online shortly after acceptance, which is prior
www.rsc.org/greenchem Volume 12 | Number 9 | September 2010 | Pages 1481–1676
to technical editing, formatting and proof reading. This free service from RSC
Publishing allows authors to make their results available to the community, in
citable form, before publication of the edited article. This Accepted Manuscript will
be replaced by the edited and formatted Advance Article as soon as this is available.

To cite this manuscript please use its permanent Digital Object Identifier (DOI®),
which is identical for all formats of publication.

More information about Accepted Manuscripts can be found in the


Information for Authors.

ISSN 1463-9262
Please note that technical editing may introduce minor changes to the text and/or
COMMUNICATION
Luque, Varma and Baruwati
Magnetically seperable organocatalyst
CRITICAL REVIEW
Dumesic et al.
Catalytic conversion of biomass
graphics contained in the manuscript submitted by the author(s) which may alter
for homocoupling of arylboronic acids to biofuels 1463-9262(2010)12:9;1-U

content, and that the standard Terms & Conditions and the ethical guidelines
that apply to the journal are still applicable. In no event shall the RSC be held
responsible for any errors or omissions in these Accepted Manuscript manuscripts or
any consequences arising from the use of any information contained in them.

www.rsc.org/greenchem
Registered Charity Number 207890
View Article Online
Page 1 of 15
Green Chemistry Green Chemistry DynamicDOI:Article Links ►
10.1039/C3GC41581C

Cite this: DOI: 10.1039/c0xx00000x


www.rsc.org/greenchem PAPER
Characteristics and origin of char and coke from fast and slow, catalytic
and thermal pyrolysis of biomass and relevant model compounds
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

Shoucheng Du, Julia Valla and George M. Bollas*


Received (in XXX, XXX) Xth XXXXXXXXX 2013, Accepted Xth XXXXXXXXX 2013
5 DOI: 10.1039/b000000x

Green Chemistry Accepted Manuscript


Char and coke from biomass catalytic pyrolysis have different origins. They cannot be lumped as one
since they occupy different locations on the catalyst surface and, thus, contribute differently to catalyst
deactivation. In this study, catalyst (ZSM-5) deactivation in the perspective of comparison of char and
coke from pyrolysis of different biomass types is investigated. Pine sawdust, glucose, and cellulose are
10 used as feedstocks in the pyrolysis experiments. Biomass char and coke samples produced via slow and
fast, thermal and catalytic pyrolysis are characterized with respect to their overall content, oxidation
reactivity, catalyst surface area, pore size distribution changes, bonding groups and their effect on catalyst
performance. In particular, it is shown that char forms as an external layer on the catalyst surface and in
its macropores, whereas coke forms inside the zeolite micropores via hydrogen transfer and addition
15 reactions. The catalyst effect on glucose and pine slow catalytic pyrolysis is minor compared with that on
cellulose slow catalytic pyrolysis, due to macropore blocking by char formation. In fast catalytic
pyrolysis, catalyst deactivation is mainly attributed to micropore blocking by coke formation. Char and
coke are shown to coexist on the catalyst surface after fast catalytic experiments, with the char content
after glucose fast catalytic pyrolysis being 30wt% of the total solid residue. The origins of char and coke
20 in the cellulose, hemicellulose and lignin components of pine are identified and mechanisms for their
formation are proposed.

Introduction condensable products and condensable bio-oil plus water.9,17,18


However, the definition of coke and char varies among different
Biomass has received considerable attention as a close to CO2
studies. Elordi et al.19 studied the fast catalytic pyrolysis of
neutral and sustainable feedstock that can replace fossil fuels for
50 polyethylene in a spouted bed reactor at 500°C, using HZSM-5,
25 energy generation.1–5 Lignocellulosic biomass is a low-cost
HY, Hβ catalysts. They defined coke as the carbonaceous
feedstock that is uniquely suited for the production of sustainable
material deposited on the catalyst. They claimed that the
liquid fuels.6,4,7 However, lignocellulosic biomass is difficult to
combustion of coke in the meso- and macro-pores of the catalyst
deconstruct into hydrocarbon-containing sub-fractions, because
shows a temperature programmed oxidation (TPO) peak at lower
of its heterogeneous composition.7 The approaches used for
55 temperatures, compared with the coke located inside the zeolite
30 lignocellulose deconstruction can be broadly lumped into
crystal channels due to differences in composition. They also
combustion, gasification, liquefaction, and pyrolysis.8,9 Among
observed coke outside the zeolite crystals with heterogeneous
these, pyrolysis is developing rapidly and can play a very
sizes between 10 and 50 nm, using transmission electron
important role in the future of renewable energy production.10
microscopy (TEM). In the review of biofuel production by Huber
Depending on the heating rate, biomass pyrolysis can be
60 and Corma,20 coke was defined as the organic fraction that could
35 separated into two categories: slow pyrolysis (0.1 1 K/s) and fast
only be removed from the catalyst via calcination. Char was
pyrolysis (10 1000 K/s or higher, including flash pyrolysis).11
defined as the organics deposited in the reactor as a result of
Slow pyrolysis produces large amounts of carbonaceous residues,
thermal decomposition, but not on the catalyst. Triantafyllidis et
which can be used as a solid fuel or fertilizer, whereas fast
al.21 studied the fast catalytic pyrolysis of beech wood in a fixed
pyrolysis produces high yields of bio-oil.6 Fast catalytic pyrolysis
65 bed reactor at 500°C, using mesoporous aluminosilicate and
40 enhances selectivity to hydrocarbons, particularly aromatics.12–15
conventional Al-MCM-41 catalysts. They considered coke as a
The reason for introducing a catalyst to the pyrolysis process is
lump of the solid carbonaceous residues produced thermally in
primarily because of its ability to improve the quality of bio-oil.16
the reactor as a separate phase to the catalyst, as well as the solid
residues deposited on the catalyst surface due to thermal and
In biomass pyrolysis, slow or fast, catalytic or thermal, there is
70 catalytic cracking. Generally, coke is considered as the catalytic
45 always some solid residue (typically, a mixture of coke and char,
product, whereas char is the residue formed via thermal
depending on the pyrolysis process) produced in parallel to non-
deconstruction. This definition is widely accepted, and many

This journal is © The Royal Society of Chemistry [year] [journal], [year], [vol], 00–00 | 1
View Article Online
Green Chemistry Page 2 of 15
DOI: 10.1039/C3GC41581C

Char & Catalyst Collection T


vent pyrolysis char and coke characteristics. The objective of this
Liquid & Gas Analysis
vent
T T study is to explore the difference between coke and char residues
and explore the origin of their formation and their contribution to
45 catalyst deactivation during biomass catalytic pyrolysis.
T

Experimental section
Reactor
Experimental setup
T
Two experimental setups are used in the pyrolysis experiments.
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

T
Liquid Collection T
Filter Slow pyrolysis is performed in a fixed bed quartz reactor (1 inch
Flow Meter
Gas Analysis 50 o.d. and 24 inch length), which is heated in a horizontal tube
FTIR real furnace. Fast pyrolysis is studied in a specially designed spouted
time analyzer bed reactor (Fig.1). In the spouted bed reactor, biomass is fed
Gas
Collection with an auger feeder at the bottom and enters the reactor via
vent

Green Chemistry Accepted Manuscript


N2 entrainment with 4-10 L/min N2 flow. The reactor operates at
55 temperatures up to 1000°C, 600°C in the experiments discussed
Coolant Auger
Feeder
here. The products pass through six impingers, with three of them
N2 filled with 5 ml methanol each, to collect the liquid products. A
Fig.1 Experiment setup for biomass fast (catalytic) pyrolysis. cooling jacket is used at the bottom of the inlet tubing of the
reactor, to prevent thermal pyrolysis at lower temperatures (or
researchers describe the solid residue after thermal pyrolysis as 60 temperature gradients) before entering the spouting zone. As
char.22–24 Based on this definition, primary decomposition and shown by Ferdous et al.,33 Nowakowski et al.,34 and Sharma et
5 secondary polymerization contribute to char formation,25,26 while al.,35 lignin can melt at lower temperatures before entering the hot
coke formation is mainly attributed to catalytic polymerization of reactor zone, which makes it very difficult to study the effect of
small biomass molecules inside catalyst pores.27 lignin on the pyrolysis of lignocellulosic biomass, as most of the
65 lignin is never fed to the reactor. This is addressed in the current
Generally, coke formation leads to catalyst deactivation and setup with the aforementioned cooling jacket, keeping the feeding
10 results in undesirable product selectivity in biomass pyrolysis,6 pipe under 100°C, and the high gas velocities in the feeding pipe,
whereas char may or may not deactivate the catalyst, depending which prohibit early decomposition of biomass components.
on the location of its formation. However, the real reason for
catalyst deactivation due to coke and char is not well understood. 70 The selection of a spouted bed reactor for biomass pyrolysis was
Carlson et al.28 studied the effect of ZSM-5 deactivation (due to based on previous work by Cui and Grace,2 Atutxa et al.16 and
15 coking) on the selectivity of glucose fast catalytic pyrolysis in a Bilbao and co-workers.36–38 Spouted bed reactors are ideal for the
pyroprobe reactor. Coke yields of the order of 33 – 45 mol% characteristics of biomass: they can handle large particle size
(moles of carbon in coke per total moles of carbon) were distributions, larger particles, differences in particle densities, and
measured, which translates to about 15 wt% coke yield on 75 provide excellent mixing.2 Their fountain-type hydrodynamic
glucose mass basis. In experiments with zeolite to glucose ratios regime can decouple residence times of gas and solids; thus,
20 of 19 (thus, coke on catalyst of about 0.79 wt%), they observed reducing unwanted secondary reactions,37 and provides high
positive effects of coke formation on biomass pyrolysis heating rates and isothermality.38 The detailed reactor
selectivity. On the contrary, Aho et al.29 performed fast catalytic dimensions, design equations and calculations of the reactor
pyrolysis experiments with pine and ZSM-5 in a fluidized bed 80 operating variables are discussed in Du and Bollas (2013).39
reactor at 450°C, but with low zeolite to biomass ratio (0.4),
25 showing that coking of the zeolite leads to a significant decrease Feedstock and catalyst
in catalytic activity. Also, Cheng and Huber13 investigated the In order to study the mechanism of biomass pyrolysis, various
conversion of furan over HZSM-5 in a fixed-bed reactor at biomass feedstocks and model compounds are being used.
600°C. A continuous loss of catalytic activity was observed as the Among them, pine sawdust is widely used in bio-fuel production
amount of coke increased on the catalyst surface. In most studies, 85 studies.40,12,38 Cellulose and glucose are popular biomass model
30 carbon and char are lumped as “coke”, but according to Aho et compounds. Cellulose is the major component in cellulosic
al.29 the char to coke ratio after pyrolysis of pine sawdust with biomass (typically, 23–32% in lignocellulose); therefore, the
ZSM-5 is about 2:1. study of cellulose pyrolysis has been considered to be critical for
the study of biomass pyrolysis mechanisms.41,42,22 Glucose, the
As indicated above, significant effort has been devoted on the 90 monomer of cellulose, is often used as a model compound for
35 characterization of biomass char and coke. However, most of the cellulose to simplify simulations.26 Recently, Mettler et al.43
published work focuses on thermal chars,30–32 leading to a lack of studied the fast thermal pyrolysis of glucose, cellodextrins and
understanding of the mechanisms and effects of char and coke cellulose in a thin-film pyrolysis reactor at 500°C. They reported
formation when catalyst is introduced in the pyrolysis. This work very different product distributions between glucose (with –OH
focuses on studying catalyst deactivation and the effectiveness of 95 groups instead of glycosidic linkages) and cellulose pyrolysis,
40 model compounds, in particular glucose and cellulose, as showing the inappropriateness of using glucose as a model
compared to pine from the perspective of comparing their compound for cellulose.

2 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society of Chemistry [year]
View Article Online
Page 3 of 15 Green Chemistry DOI: 10.1039/C3GC41581C

Particle milled on “first side.” Particle milled on “second side.” Coupon “lifted out” using
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

Coupon imaged in STEM


Cross sectioned on equator (View from above) Omniprobe micro manipulator
Fig.2 Process of sample preparation in Focused ion beam (FIB)/Energy Dispersive X-Ray Spectroscopy (EDX) analysis.

According to the above discussion, pine, cellulose and glucose experiments, assuming zero catalytic activity.
were chosen as the feedstocks in this study. Sawdust from pine
Char/coke characterization
bark (50.81wt% C, 5.95wt% H, 42.96wt% O, 0.28wt% N), α-

Green Chemistry Accepted Manuscript


5 cellulose from Sigma-Aldrich and α-D-glucose from Acros 50 Scanning Electron Microscopy (SEM)
Organics were used as biomass and corresponding model Char/coke samples were observed using a FEI Quanta FEG 250
compounds. The grinded pine sawdust and glucose and cellulose scanning electron microscopy (SEM) under high vacuum, to
powders were sieved to <350 μm in particle size. distinguish the differences in morphology of char samples
produced in different experiments and visualize the coke/char
10 In biomass catalytic pyrolysis, zeolite catalysts, including ZSM-5, 55 deposition on the surface of catalyst pellets. Before each
Beta zeolite, Y zeolite, Mordenite and several mesoporous experiment, char samples were coated with gold to inhibit
materials, have been widely studied.44,29,45,21 In the majority of charging when the magnification is high.
these studies, ZSM-5 has been proven to be very effective in Focused Ion Beam (FIB)/Energy Dispersive X-Ray
Spectroscopy (EDX) analysis
catalytic (fast) pyrolysis due to its proper pore morphology. Jae et
60 The FIB in-situ sample preparation and EDX element mapping
15 al.44 investigated the shape selectivity of zeolite catalysts for
was performed in a FEI Strata 400 STEM Dual Beam system, a
glucose conversion in fast catalytic pyrolysis in a pyroprobe
fully digital Field Emission Scanning Electron Microscope (FE-
reactor at 600°C, using different zeolite catalysts. They showed
SEM) equipped with Focused Ion Beam (FIB) technology and
that medium pore zeolites, such as ZSM-5 had the highest
Flip stage/STEM assembly. A brief description of the sample
aromatic yield and minimum coke formation. Similarly, Carlson
65 preparation is illustrated in Fig.2. A catalyst particle was selected
20 et al.46 tested different zeolite catalysts for the conversion of
and milled on both sides, leaving a thin layer along the equatorial
glucose, xylitol, cellobiose and cellulose to aromatics using a
plane of the particle. The thin catalyst layer was lifted out with an
pyroprobe reactor at 600°C. They found that ZSM-5 had the
Omniprobe micro-manipulator and imaged under scanning
highest aromatic yields and the lowest coke selectivity. Thus, in
transmission electron microscopy (STEM).
this study, a commercial ZSM-5 catalyst (synthesized by W.R.
70 Thermal Gravimetric Analysis (TGA)
25 Grace & Co. in a macroporous matrix of mean particle size of 75
Comparison of thermal properties of char samples was performed
μm) was used for the catalytic studies.
in a Q-500 thermogravimetric analyzer from TA Instruments.
Experimental procedure Platinum crucibles instead of alumina crucibles were used for the
oxidation of chars, to prevent catalyst sintering with the crucibles.
All experiments presented in this work were performed with a
75 The samples were first dried at 120°C for 30 min and heated up
biomass to catalyst weight ratio of 1. The relatively low ratio was
to 900°C in air flow of 60 ml/min. Miura and Silveston47 studied
30 chosen to maximize char/coke to catalyst ratios, and thus,
non-catalytic gas-solid reactions using the temperature-
exemplify the results. In the slow catalytic pyrolysis experiments,
programmed reaction (TPR) technique. They showed a relative
biomass and catalyst were well mixed before the experiment.
unreliability of rate parameters obtained based on TPR at only
Then, the mixture was pyrolyzed in a N2 environment. Before
80 one heating rate. Therefore, in this study three different heating
starting the experiment, N2 flow (20 ml/min) was kept for 1 hr to
rates (5/10/15 K/min) were used in order to get accurate results in
35 purge the air inside the reactor. The feedstock was first dried at
TPR data processing. All the TGA results presented in this paper
120°C for 30 min and then the temperature was ramped to 600°C
were normalized to char weight loss (catalyst weight excluded) to
at a rate of 10 K/min. In fast pyrolysis, 1-2 gr of biomass was
compare thermal with catalytic residues.
dried at 120°C overnight in a separate furnace. The desired 85 Nitrogen Adsorption
reactor temperature (600°C) was reached before feeding the The surface area and pore size distribution of catalyst and coked
40 catalyst and biomass. Catalyst was fed via entrainment from the catalyst/char mixtures were determined in a Micrometitics ASAP
bottom of the reactor with 4 L/min N2 flow. When the catalyst 2020 Accelerated Surface Area and Porosimetry System. The
bed temperature reached 600°C, biomass was fed from the isotherms of N2 at 77 K were obtained from physisorption.
bottom as well. With the high N2 flow rate, biomass reaches the 90 Before analysis, all the char/catalyst samples were degassed at
spouting zone within very short times (~0.1 ms) and contacts with 250°C under vacuum for 12 hr to remove the surface
45 the hot catalyst particles. After the reaction, char/coke samples contaminants. The pore size distribution of the char/catalyst
were collected via entrainment with 19 L/min N2 flow. Quartz samples was determined from the N2 adsorption isotherms at
beads (sieved <180 μm) were used for the fast thermal pyrolysis 77 K, using the Barrett-Joyner-Halenda (BJH) method.

This journal is © The Royal Society of Chemistry [year] Journal Name, [year], [vol], 00–00 | 3
View Article Online
Green Chemistry Page 4 of 15
DOI: 10.1039/C3GC41581C

Fourier Transform Infrared Spectroscopy (FTIR)


FTIR measurements to identify functional groups in char/coke
were performed in a Nicolet MAGNA-IR 560 spectrometer with
a DTGS detector, operated at 4 cm-1 resolution and 132 scans.
5 Raman Spectroscopy
Raman spectra were obtained in a Renishaw 2000 Ramanscope,
operated with a 514 nm laser excitation source, at 1-25% power
and 16-32 exposure times, to avoid detector saturation. For each
sample, laser focus was set to 40% to prevent local damage and 3
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

10 different positions were analysed to verify the spectra.


Table 1 Comparison of char/coke yields between slow and fast, thermal
and catalytic pyrolysis

Char/coke yield wt% Glucose Cellulose Pine


32.6752

Green Chemistry Accepted Manuscript


1048 18.7-24.449
48 1049 29.653
Slow thermal (Literature) 18
750 24-2754
20.151 29.5-32.655
21.551
Slow thermal (this work) 20.1 16.1 31.2
Slow catalytic (Literature) 2356 - 19.6-27.857
Slow catalytic (this work) 22.9 24.6 35.7
23.743
Fast thermal (Literature) 5.3558 22.616
9.8458
Fast thermal (this work) 7.6 8.7 14.9
1528 17.6-18.816
Fast catalytic (Literature) 1246
1446 14.259
Fast catalytic (this work) 8.7 9.3 15.5

Results and discussion


15 Char/coke yields
Table 1 presents a brief literature review and experimental results 45 Fig.3 SEM morphology of biomass feedstocks: (a) glucose, (b) cellulose,
(c) pine; and chars obtained by: (d) glucose slow thermal, (e) cellulose
of the current study for the yields of char/coke from pyrolysis of
slow thermal, (f) pine slow thermal, (g) glucose slow catalytic, (h)
glucose, cellulose and pine. The effect of heating rates and cellulose slow catalytic, (i) pine slow catalytic, (j) glucose fast thermal,
catalyst are investigated in this study, although other factors, such (k) cellulose fast thermal, (l) pine fast thermal, (m) glucose fast catalytic,
20 as temperature, also play an important role 31,32,17. The char/coke 50 (n) cellulose fast catalytic, (o) pine fast catalytic pyrolysis.
yields are significantly decreasing when fast heating rates are
Char/coke morphologies
applied to the pyrolysis compared with slow heating rates. The
lower char/coke yields from fast heating rates can be explained In Fig.3 the surface morphological characteristics of pure
by the enhancement of the bond-scission reactions of the biomass biomass feedstocks (a-c) and corresponding chars with or without
25 to form tar fragments, which, to some extent, limits the secondary catalyst, collected after slow thermal, (d-f), slow catalytic, (g-i),
pyrolysis (polymerization) of the volatiles.60 Moreover, char/coke 55 fast thermal, (j-l), and fast catalytic, (m-o), pyrolysis of the three
yields vary significantly between different feedstocks. To be types of biomass are investigated using SEM. The morphology of
specific, in slow thermal pyrolysis, pine produces the highest char derived from the slow thermal pyrolysis of glucose, (d),
yield of char/coke; whereas cellulose produces less char/coke cellulose, (e), and pine, (f), retains a similarity to the original
30 than glucose. Similar results were also obtained by other structure of the feedstock (a-c). In the case of slow catalytic
researchers.58,61 In slow catalytic pyrolysis, fast thermal pyrolysis 60 pyrolysis, the glucose char, (g), is surrounding the catalyst
and fast catalytic pyrolysis, char/coke yield from pyrolysis of particles forming catalyst agglomerates due to low temperature
pine outweighs that from cellulose, while glucose produces the primary pyrolysis and melting. The slow catalytic cellulose char,
lowest yield of char/coke. Furthermore, in order to study the (h), has a spiral type structure similar to the respective char
35 catalyst effect, char/coke yields between thermal pyrolysis and produced from slow thermal pyrolysis, and appears to be formed
catalytic pyrolysis are compared. For all the pyrolysis conditions 65 as a separate phase to the catalyst particles. Pine slow catalytic
performed in this study, catalytic pyrolysis produces more char, (i), consist of irregular, large particles with slit-shaped
carbonaceous residues than the corresponding thermal pyrolysis surfaces surrounding the catalyst particles. In fast thermal
due to the presence of catalyst. However, the reason for the pyrolysis, the morphologies of chars from glucose, (j), cellulose,
40 increase of carbonaceous solid residues is not certain at this point. (k), and pine, (l), retain the original structure to some extent, but
In other words, the solid residues can be char (a non-catalytic 70 in a different way to the slow thermal pyrolysis, due to the
product), catalytic coke, or both. In the following, experimental different heating rates. A significant difference in morphology
results are analysed, focusing on the formation of char and coke. can be seen in pine fast thermal char, (l), compared with the slow
thermal pyrolysis.

4 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society of Chemistry [year]
View Article Online
Page 5 of 15 Green Chemistry DOI: 10.1039/C3GC41581C

no edges and only some small red spots can be seen, which
indicates the formation of carbon inside the zeolite. In all the fast
catalytic pyrolysis cases, the catalyst has char deposition both on
the outer surface of the catalyst and inside the catalyst.
50 Specifically, in fast catalytic pyrolysis of glucose, there exists a
thinner and more homogeneous carbon layer on the surface of the
catalyst pellet compared with slow catalytic pyrolysis of glucose,
which verifies the improvement in catalyst accessibility,
concluded by the SEM analysis. Overall, the carbon mapping of
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

55 the coked catalyst shows significant differences between slow


catalytic pyrolysis of glucose/sawdust and cellulose. The
deposition of char on the outer surface of the catalyst might affect
the accessibility of the catalyst at the beginning of the pyrolysis.
Thus, it can prevent volatile matters from accessing the zeolite
Fig.4 FIB images of the equatorial plane of the catalyst particle after slow

Green Chemistry Accepted Manuscript


60 and prohibit secondary pyrolysis, where coke is produced.
and fast catalytic pyrolysis. (a) glucose slow catalytic, (b) cellulose slow
catalytic, (c) pine slow catalytic, (d) glucose fast catalytic, (e) cellulose Table 2 Micropore and total surface area of the fresh ZSM-5 catalyst and
5 fast catalytic, (f) pine fast catalytic. (Green=Al, Yellow=Si, Red=Carbon) coked catalyst/char mixtures after catalytic pyrolysis (m2/g)
The fast volatile release during fast pyrolysis enlarges the internal Micropore area (<2nm) Total surface area
cavities, resulting in a more open structure. Therefore, the Pure catalyst 98.98 124.3
macroporosity of chars increases with increasing heating rate.17 Glucose slow catalytic 122.4 162.1
Glucose fast catalytic 47.55 68.33
In fast catalytic pyrolysis, there is no catalyst agglomeration Cellulose slow catalytic 86.30 115.0
10 occurring in the glucose case, (m), reflecting the benefit of fast Cellulose fast catalytic 60.93 90.05
heating rates to prevent excessive char formation. In the case of Pine slow catalytic 109.7 141.5
fast catalytic pyrolysis of cellulose, (n), and pine, (o), char Pine fast catalytic 76.94 99.50
generally does not retain the structure of the original feedstock
Comparison of surface areas and pore size distributions
indicating that repolymerization may be the dominant mechanism
15 for its formation.62 However, it should be noted that in all the 65 Fig.5 shows the isotherms of N2 adsorption and corresponding
pyrolysis experiments, the formation of coke is considered to pore size distribution plots for pure ZSM-5 catalyst and the solid
occur inside the catalyst pores due to catalytically enhanced residuals obtained after slow and fast catalytic pyrolysis. The
reactions of small molecule products and intermediates in the isotherm profile of the catalyst after glucose slow catalytic
secondary pyrolysis of volatile matters.62 Thus, the formation of pyrolysis exhibits very characteristic profiles. Microporosity (<2
20 coke cannot be visualized with SEM. 70 nm) increases, which can be attributed to the porosity of char
formed, also evident in the desorption profiles that indicate ink
The SEM results imply that the catalyst effect on pyrolysis is bottle shaped pores. Moreover, the catalyst macropores are
different for the biomass feedstocks studied, as far as formation significantly diminished, while the hysteresis loop during N2
of char on the outer surface of catalyst particles is concerned. desorption reflects the creation of a mesoporous network in the
25 Evidently, in slow catalytic pyrolysis, glucose and pine first melt, 75 form of ink bottle type pores (hysteresis type H2 in IUPAC
wetting the catalyst surface, and then pyrolyze, leaving classification), as shown in the isotherm and pore size distribution
particle/char aggregates. It is reasonable to assume that the plots. In the isotherm of the solid residue after cellulose slow
catalyst quickly deactivates in these conditions, by losing its catalytic pyrolysis, the decrease of the micropore volume reflects
accessibility due to the rapid surface coverage of a liquid the enhanced coke formation and micropore blocking due to
30 pyrolysis intermediate. Therefore, only minimal differences 80 coking. In the isotherm of the solid residue after pine slow
between the characteristics of the char/catalyst mixtures after catalytic pyrolysis, the micropore volume increases slightly, but
slow pyrolysis of glucose and pine should be expected. the macropore volume does not have a significant change (shown
also in the pore size distribution plot). In fast catalytic pyrolysis,
Char/coke deposition on the outer surface and the equatorial
plane of the catalyst however, a clear loss of microporosity is observed in the
85 isotherms for all the cases, which can be attributed to the
35 Fig.4 is the EDX elemental mapping of coked catalysts from formation of coke inside the zeolite micropores. Table 2 shows a
pyrolysis of different biomass feedstocks. Clear char “footprints” summary of the micropore area and total surface area of the
can be seen in the slow catalytic pyrolysis of glucose and catalyst and the catalytic chars (catalyst included). The micropore
sawdust. The catalyst coupon from slow catalytic pyrolysis of area of glucose and pine slow catalytic chars increased by 24%
glucose has a clear char edge, which is consistent with SEM 90 and 11% respectively, compared to pure catalyst. The
40 pictures. It also has a large amount of char inside the pellet, microporosity of cellulose slow catalytic chars decreased by 13%,
which must be the result of melting inside catalyst macropores at a smaller amount compared with the microporosity of all the fast
intermediate temperatures. The coupon from slow catalytic catalytic chars, which decreased by 52% (glucose), 38%
pyrolysis of pine also has a char edge, but it is much thinner than (cellulose), and 22% (pine). The glucose slow catalytic char has
that of glucose. Inside the catalyst pellet, less carbon is observed. 95 the highest total surface area (highest microporosity), whereas the
45 The catalyst after slow catalytic pyrolysis of cellulose has almost glucose fast catalytic char has the lowest microporosity.

This journal is © The Royal Society of Chemistry [year] Journal Name, [year], [vol], 00–00 | 5
View Article Online
Green Chemistry DOI: 10.1039/C3GC41581C
Page 6 of 15

 r rsmnrG b(d b(d b(d

 r rsmnrG

 r rsmnrG


r  r  r 
bdd ²  r r  bdd .  r r  bdd r r 
²  r r  .  r r  r r 
Td Td Td

gd gd gd

3d 3d 3d
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

(d (d (d

d d d
d dQ( dQ3 dQg dQT b d dQ( dQ3 dQg dQT b d dQ( dQ3 dQg dQT b
D # $r% rsn G E # $r% rsn G F # $r% rsn G
/d /d bmd
3/ 3/ b(d
vn sCGr rvrs*n+,G

vn sCGr rvrs*n+,G

vn sCGr rvrs*n+,G


3d r  3d r  bbd r 

Green Chemistry Accepted Manuscript


²  r r  .  r r  r r 
m/ ²  r r  m/ .  r r  bdd r r 
md md )d
bd bd bd

T T T

g g g

3 3 3

( ( (

d d d
d /d bdd b/d (dd (/d d /d bdd b/d (dd (/d d /d bdd b/d (dd (/d
G  r .r  rsG H  r .r  rsG I  r .r  rsG
Fig.5 N2 adsorption/desorption isotherms measured at −196°C and the pore size distributions of pure catalyst and deactivated catalyst produced from (a,d)
glucose, (b,e) cellulose, (c,f) pine, calculated from adsorption isotherms by using the BJH method.

This result reveals that the catalyst porosity change is most for the glucose thermal char and at ~500°C for the cellulose and
5 significant in the pyrolysis of glucose when high heating rate is 35 pine thermal char. It has been documented in the literature62 that
introduced in the experiment. Combining the results from SEM, glucose char is produced via polymerization reactions, whereas
FIB-STEM, BET and TPO (discussed in the next section), the pyrolytic decomposition is the main pathway for cellulose char
significant reduction of macropores in the case of glucose and formation. Evidently, these two different mechanisms are
pine slow catalytic char might be attributed to macropore responsible for different char structures and/or char composition,
10 blocking by char formation on the catalyst outer surface. 40 which correspondingly react at different temperatures.19 The char
Macropore blockage is the reason for the minor catalyst effect on formed from pine thermal pyrolysis exhibits a DTG maximum
the properties of char and coke (e.g., oxygen content) in slow identical to that of cellulose, whereas it also gives two additional
catalytic pyrolysis, which are responsible for TPO profiles similar DTG peaks. The three DTG peaks of pine thermal pyrolysis char
with those of slow pyrolysis char. Furthermore, the decrease of reflect char structures of different origins, corresponding to its
15 the number of micropores after cellulose slow catalytic pyrolysis 45 hemicellulose, cellulose and lignin compounds.64
is attributed to coking, showing catalytic activity, even at slow
heating rates. The smaller quantity of char produced in this case Interestingly, the addition of catalyst in slow pyrolysis conditions
is because of the catalytic enhancement of depolymerization of has only minor effect on the oxidation profiles of the solid
biomass to liquid and gas products. residues from glucose and pine. In the case of glucose, the DTG
50 peak is shifted slightly to the left, indicating a slight catalytic
20 Thermal gravimetric analysis of char/coke
activity affecting the polymerization reactions, responsible for
In order to verify the effect of char/coke formation on the catalyst glucose char/coke formation. Overall and in agreement with the
performance, temperature programmed oxidation (TPO) of the SEM images, the catalytic effect in slow pyrolysis conditions is
char/coke samples are performed in air using 3 different heating minor, which can be attributed to the catalyst surface coverage by
rates.63 Before analysing the char/coke samples, the catalyst 55 the intermediate liquid and, therefore, rapid catalyst deactivation.
25 effect on the oxidation itself was studied. TPOs of pure graphitic Small catalyst effect in slow pyrolysis of pine has also been
carbon and carbon/catalyst mixtures showed that there is no observed in the literature,65 where the pyrolysis of wood was
significant effect of the catalyst on the char/coke oxidation. Fig.6 performed in TGA. Cellulose pyrolysis exhibits a rather different
presents differential weight loss profiles in TPO of the solid behaviour. The DTG peak moves to the right, indicating a more
residues of thermal and catalytic (slow and fast) pyrolysis. TPO 60 significant catalytic activity towards solid residues of lower
30 profiles are normalized, to exclude the catalyst weight and the oxygen content. This indicates a formation of an oxygen poorer
first derivative of weight loss (DTG) versus temperature is phase that can be attributed to coke formation or formation of
shown. In the comparison of slow thermal and slow catalytic gaseous intermediates that react with the cellulose residues
char/coke from the three feedstocks, the DTG peak is at ~550°C forming lower oxygen content char. Moreover, the oxidation peak

6 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society of Chemistry [year]
View Article Online
Page 7 of 15 Green Chemistry DOI: 10.1039/C3GC41581C

T0 10 the cellulose char/coke oxidation.66 TPO of fast catalytic


D o o0 p io oyo0TT%
o oTk p io oyo017%
o oT0 p io oyo00c% pyrolysis samples show a broad DTG peak for glucose and
o  o0 p io oyo0k0 %
u o  oTk p io oyo0:c %
multiple DTG peaks for the cellulose and pine solid residues. The
T:,0 o  oT0 p io oyo0C3 %
u "o o0 p io oyoC7C%
u "o oTk p io oyo0k:%
u "o oT0 p io oyo0T3%
broad DTG peak in the case of glucose can be deconvoluted into
u "o  o0 p io oyo0k:%
u "o  oTk p io oyo0:7 % multiple peaks that represent different oxidation steps, consistent
u "o  oT0 p io oyo0CC %
Tk 15 with cellulose and pine. Generally, the first char/coke oxidation
peak at 400-450°C corresponds to char formation due to thermal
)8oaep 5

decomposition. The second DTG peak at 550-600°C corresponds


to catalytic coke formation. The wider shape of the DTG curves
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

in the case of fast catalytic pyrolysis char/coke indicates a lower


0 20 char/coke apparent oxidation reaction rate or a multiplicity of
compounds, being oxidized at different rates. Compared with fast
:,0
catalytic pyrolysis, fast thermal pyrolysis produces higher oxygen
content char, leading to a lower temperature peak in the case of

Green Chemistry Accepted Manuscript


glucose and pine. Moreover, there is no secondary DTG peak at
k
:kk 1kk Ckk 0kk 3kk 7kk 9kk ckk 25 higher temperature in the fast thermal pyrolysis, reflecting the
  ) oa%5 infeasibility of coke formation due to non-catalytic reactions. It is
1k
clearly evident that the catalytic effect is much stronger in fast
E o o0 p io oyoC37%
o oTk p io oyoC93%
o oT0 p io oyoCc9% pyrolysis experiments for each biomass feedstock compared with
o  o0 p io oyoCcc %
u o  oTk p io oyo0:: %
that in slow pyrolysis experiments.
:0 o  oT0 p io oyo019 %
u "o  o0 p io oyo010%
u "o  oTk p io oyo03: % 30
u "o  oT0 p io oyo0c: %
In summary, many hypotheses derived from SEM, STEM and
:k BET were verified with TPO results. Cellulose exhibits an
entirely different behaviour, showing smaller catalyst surface
)8oaep 5

T0 coverage at slow pyrolysis conditions. TPO confirms this


35 observation, showing a clear contribution of the catalyst to the
oxidation temperature of cellulose slow catalytic char. The
Tk catalyst/char samples after fast catalytic pyrolysis of all the
feedstocks show smaller amounts of char and intermediate
0 surface coverage of the catalyst, indicating a catalytic mechanism
40 for coke formation on the catalyst surface.
Deconvolution of DTG thermogravimetric measurements
k
:kk 1kk Ckk 0kk 3kk 7kk 9kk ckk
  ) oa%5 The TPO results of Fig.6 were further analysed using model-
T: based and statistical deconvolution methods. For this illustration
F o o0 p io oyoC3C %
o oTk p io oyoC90%
o oT0 p io oyo0kT%


the TPO experiments of the glucose char/coke and the chars of


o  o0 p io oyoC3C %
u o  oTk p io oyoC9c % 45 the thermal pyrolysis of pine were analysed. In principle, glucose
Tk o  oT0 p io oyo0k: %
u "o o0 p io oyoCT7%
u "o oTk p io oyoCC3%
u "o oT0 p io oyoC3C%
is an ideal candidate for a model-based analysis of its coke TPO,
u "o  o0 p io oyoCC:%
u "o  oTk p io oyoC7: % since there exists mainly one mechanism (i.e., polymerization)
u "o  oT0 p io oyoC9T %
9 for its formation.62 The random pore model (RPM) by Bhatia and
Perlmutter67 was utilized for this analysis. This model considers
)8oaep 5

3 50 the overlapping of pore surfaces and the competing effects of


pore growth during gasification, and the destruction of the pores
due to the coalescence of neighbouring pores by oxidation. The
C RPM models solid conversion, , according to:

( )√( ( ), (1)
:

55 where is the reaction rate constant and a pore structure


k factor of the unreacted sample:
:kk 1kk Ckk 0kk 3kk 7kk 9kk ckk
  ) oa%5
, (2)
Fig.6 Oxidation of slow thermal (solid line) slow catalytic (dotted line),
fast thermal (solid disc) and fast catalytic (empty disc) pyrolysis ( )
char/coke from (a) glucose, (b) cellulose and (c) pine in TGA in three , (3)
5 heating rates (5K/min (green), 10K/min (blue), 15K/min (red)).
of the glucose (catalytic and thermal) char/coke sample is more with , and the pore surface area, pore length, and solid
symmetrical than that of cellulose. In the latter case the high 60 porosity, respectively. Table 3 summarizes the parameters of the
temperature side of the DTG profile displays a faster decrease in best fit of the RPM for the char/coke obtained from glucose
reaction rate, which can be attributed to a lower reaction order for pyrolysis. As shown in Fig.7, the RPM is capable of representing

This journal is © The Royal Society of Chemistry [year] Journal Name, [year], [vol], 00–00 | 7
View Article Online
Green Chemistry DOI: 10.1039/C3GC41581C
Page 8 of 15

P4 D PM1. D
Pt o o2 0 
o oPR 0 
6s  s. 0 
6s  sPR 0 
o oP2 0  PR 6s  sP. 0 
PM oo2 0  ss. 0 
ooPR 0  s sPR 0 
PR ooP2 0  C1. ssP. 0 

8
4 .
t M1.
M
R R
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

P4 E C1. E
Pt o○  o2 0 
o○  oPR 0 
6s○ s. 0 
6s○ sPR 0 
o○  oP2 0  6s○ sP. 0 
PM oo2 0  ss. 0 
3 oae0 5

38sle0 5
ooPR 0  . ssPR 0 
PR ooP2 0  ssP. 0 

8
4

Green Chemistry Accepted Manuscript


M1.
t
M
R R
P4 F C1. F
Pt o○  o2 0 
o○  oPR 0 
6s○ s. 0 
6s○ sPR 0 
PM o○  oP2 0  "# 6s○ sP. 0 
oo2 0  ss. 0 
ooPR 0  . ssPR 0 
PR ooP2 0  ssP. 0 

8 
4 M1.
t
M
R R
MRR fRR tRR 2RR 4RR 6RR 8RR CRR MRR fRR 2RR .RR 7RR CRR yRR cRR
 $ % oa&○5  $ % sl&○5
Fig.7 Application of the RPM on the TPO of chars from: (a) glucose slow Fig.8 Application of the RPM on the TPO of chars from: (a) glucose fast
thermal pyrolysis; (b) glucose slow catalytic pyrolysis; and (c) glucose thermal pyrolysis; (b) glucose fast catalytic pyrolysis; and (c) glucose fast
slow catalytic pyrolysis using the RPM parameters of (a). 25 catalytic pyrolysis using the combined RPM models of (a) and (b).

5 the TPO of chars from glucose slow pyrolysis. The small Fig.8 presents the RPM fit of the glucose chars from fast
factors estimated should be attributed to the high surface area of pyrolysis. The quality of fit is again very good for the thermal
glucose chars and the presumably small pore length (indicated by pyrolysis char, whereas the catalytic pyrolysis char exhibits two
the EDX elemental mapping of Fig.4). convoluted DTG peaks that cannot be represented by the RPM
30 (Fig.8(b)). Fig.8(c) presents the result of the RPM when including
Table 3 Kinetic parameters of the RPM for the glucose char samples
10 during TPO at three heating rates (5, 10 and 15 K min −1) the contribution of the fast thermal char, adjusting only the RPM
shape factor, to account for the effect of the catalyst on shape and
(s-1) (J mol-1) R2 surface area differences. For a more detailed discussion on the
Glucose slow thermal 1.870E+6 1.361E+5 0.268 9.918E-1 application of the RPM to multicomponent solids refer to the
Glucose slow catalytic
a 8.302E+5 1.292E+5 0 9.935E-1 35 works by Miura et al.68 and Fermoso et al.63 The inclusion of the
Glucose slow catalytic RPM of the fast thermal char of glucose is capable of
b 1.870E+6 1.361E+5 0.193 9.621E-1 deconvoluting the fast catalytic DTG, identifying two solid
Glucose fast thermal 3.665E+2 0.768E+5 0.898 9.956E-1 residues of different origins. Therefore, it is reasonable to assume
Glucose fast catalytic that, in parallel to the thermal polymerization reactions (char),
c 8.066 5.932E+5 0 9.500E-1
40 there is a significant contribution from catalytically enhanced
Glucose fast catalytic
d 6.525E+1 0.755E+5 0 9.968E-1 reactions yielding solid residues of different composition (coke).
a
This is an interesting finding, because it indicates that we can
Assumes a dominant contribution of catalytic reactions; b Assumes a
negligible contribution of catalytic reactions; c Assumes only one DTG
deconvolute experimentally the extent to which the catalyst
peak; d Includes the DTG contribution of the thermal char. contributes to catalytic pyrolysis reactions. The RPM predicts
45 that 30wt% of the solid residue after fast catalytic pyrolysis of
In agreement with the previous discussion about the lack of
glucose is fast thermal char. It should be noted that application of
15 catalytic contribution in glucose slow pyrolysis, the RPM is
the RPM is superior to statistical deconvolution methods, as
capable of modelling the DTG profile of the glucose slow
kinetic constants are extracted and the deconvolution is
catalytic char with the parameters of the thermal pyrolysis fit, by
performed with the same constants for different TPO heating
only adjusting the RPM shape factor (Fig.7(c)). This clearly
50 rates. This is valid for as long as the fits are of statistical
indicates that the only contribution of the catalyst at slow
significance, which is the case when fitting a small number of
20 pyrolysis conditions is on the shape, surface area and pore length
DTG peaks (small number of solid components being oxidized).
of the char formed, but not on its chemical composition.

8 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society of Chemistry [year]
View Article Online
Page 9 of 15 Green Chemistry DOI: 10.1039/C3GC41581C

D 7 TCTTTctTT
E TCTTTctTT
F
T
G T.TTT39543TTTTT643GT
T .9 TCTTTctTT T
T.TTT3654%TTTTT6433T T.TTT32G4.TTTTT643DT
3 T9TTT33%4%TTTTT6423T T9TTT3DD42TTTTT6426T
.6 T9TTT3%94GTTTTT6426T
T2TTT3DD42TTTTT649.T T2TTT7664GTTTTT649.T T2TTT79349TTTTT649.T
T3TTT75243TTTTT646.T 5 T3TTT7G54%TTTTT646.T T3TTT56542TTTTT646.T
TcF0 t

TcF0 t

TcF0 t
2 G

3 5
9
3
. 9
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

9
6 6 6
966 266 366 766 566 D66 966 266 366 766 566 D66 966 266 366 766 566 D66
  ! Tct   ! Tct   ! Tct
G D TCTTTctTT T
H .6 TCTTTctTT T
I TCTTTctTT T
.9

Green Chemistry Accepted Manuscript


5 T.TTT2G34DTTTTT6453T T.TTT36G45TTTTT64D2T T.TTT39547TTTTT64D%T
T9TTT32249TTTTT6495T G T9TTT37D42TTTTT64.5T
.6 T9TTT3D74.TTTTT64.2T
7 T2TTT37G4DTTTTT646GT T2TTT3G247TTTTT646%T T2TTT76.42TTTTT646DT
T3TTT7294DTTTTT6469T T3TTT77.46TTTTT6469T T3TTT7534GTTTTT646.T
TcF0 t

TcF0 t

TcF0 t
5 G
3
2 5
3
9 3
9
. 9
6 6 6
966 266 366 766 566 D66 966 266 366 766 566 D66 966 266 366 766 566 D66
  ! Tct   ! Tct   ! Tct
Fig.9 Exponentially Modified Gaussian peak deconvolution of chars obtained from: (a-c) pine slow thermal pyrolysis, 5/10/15 K min-1; and (d-f) pine fast
thermal pyrolysis, 5/10/15 K min-1.

However, chars from thermal pyrolysis of pine show four 15-25%). Moreover, in fast pyrolysis of pine, the fraction of char
5 inflection points in their DTG curves, with additional peaks 35 from hemicellulose increased significantly, whereas the fraction
measurable after catalytic pyrolysis, due to the formation of of char from cellulose decreased, which reflects that fast heating
catalytic coke. Therefore, a more conventional statistical rates favour the formation of hemicellulose char. The ratio of
deconvolution approach was employed; using iterative least- cellulose to lignin char is about 6/10 in all cases, which is in good
square fits of Exponentially Modified Gaussian (EMG) agreement with their initial fractions (~40wt% and ~20wt%,
10 distribution function to the pine DTG signals. Application of 40 respectively) times their char yield (~15wt% and ~50wt%,71
EMG is very common in deconvolution of chromatographic respectively). Fig.9 shows that in the research of catalytic options
peaks69 and was selected for this analysis due to its ability to for the minimization of char formation in biomass pyrolysis, we
represent fronted (or tailed) distributions. Fronted distributions need to focus on the reaction pathways of hemicellulose, since it
are evident in all the TPOs of Fig.6, particularly in the TPO of appears to be the least affected by the fast heating rates. The char
15 cellulose. Fig.9 shows the results of the deconvolution for the 45 fractions of lignin and cellulose (peaks #2 and #3, respectively)
slow and fast pyrolysis chars from pine. For each case, the four are decreasing with increasing pyrolysis heating rates.
deconvoluted peaks represent different chars from the three main
FTIR and Raman analysis of char and coke
components in pine (hemicellulose, cellulose and lignin), which
means that these precursors participate in different reactions (or The char/coke chemical composition with respect to their
20 same type of reactions, but with different precursors). The last bonding groups was studied in FTIR. Fig.10 shows the spectrum
small peak at T>500°C is attributed to the oxidation of extractives 50 1500-1800 cm-1 of the glucose char/coke, produced at different
and other heterogeneous components of pine and should be heating rates with and without catalyst. The C=C stretch (1620-
considered part of the lignin char. These high temperature peaks 1680 cm-1) appears in the char/coke produced from all the
were evident in TPO of the original pine (not shown here); pyrolysis experiments. Interestingly, the char/coke after fast
25 therefore, they are not a pyrolysis product. By comparing the catalytic pyrolysis contains minor amounts of carbonyl (C=O)
deconvolution results with those from TPO of pine and cellulose, 55 groups, which reflects the efficiency of oxygen removal in fast
it is reasonable to assume that peaks #1, #2, #3 represent chars catalytic pyrolysis and the dominance of a different mechanism
from hemicellulose, lignin and cellulose, respectively; which is producing the catalytic coke. Consistent with the TPO results, the
consistent with the location of peaks of the three components in fast heating rates of the spouted bed reactor enhance the catalytic
30 the deconvolution of pyrolysis of pine.70 In slow pyrolysis of effect in pyrolysis, thus favouring catalytic coke formation. The
pine, the ratio of hemicellulose (peak #1) and cellulose (peak #3) 60 diminished C=O in fast catalytic pyrolysis compared to that in
chars is ~2, which does not agree well with the initial pine slow catalytic pyrolysis reveals that C=O is contained mostly in
composition (hemicellulose: 23-32%; cellulose: 38-50%; lignin: char instead of coke.

This journal is © The Royal Society of Chemistry [year] Journal Name, [year], [vol], 00–00 | 9
View Article Online
Green Chemistry Page 10 of 15
DOI: 10.1039/C3GC41581C

o   o  catalytic pyrolysis char/coke. The fraction of the areas of the G
o     o    and Gꞌ bands exhibit similar trends to those observed from TPO,
indicating an increase in catalytic coke. Furthermore, the total
    oa CCc

peak area of the defect band changes only slightly between slow
50 thermal and slow catalytic pyrolysis, compared with that between
0 fast thermal and fast catalytic pyrolysis, which is very consistent
to the TPO results (lack of significant catalytic activity at slow
0 pyrolysis conditions). It is not reasonable to define coke as the
well-structured aromatics and char as defect aromatics, since G
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

55 and D bands coexist in all the Raman spectra, but the changes of
their area fractions are indicative of the same observations in
TPO, STEM/FIB and FTIR.
TuOO TuuO T)OO T)uO TWOO TWuO TvOO Reaction pathways for coke and char formation
sT
   oa c

Green Chemistry Accepted Manuscript


Fig.10 FTIR spectra of glucose chars from slow thermal pyrolysis, slow Dominant mechanisms for the formation of coke and char in fast
catalytic pyrolysis, fast thermal pyrolysis and fast catalytic pyrolysis 60 catalytic pyrolysis of glucose28 and fast thermal pyrolysis of
cellulose41 have been postulated in the literature. In summary, the
The spectrum at 2500-3600 cm-1 (not shown here), exhibits a formation of char and coke can be described as the result of
5 similar broad peak for all the chars studied, showing that O-H polymerization, dehydration, decarboxylation, decarbonylation of
bonds (free hydroxyl bonded, hydrogen bonded and in carboxyl three origins: (1) anhydrosugars, (2) furanic compounds, (3)
group) exist consistently. The FTIR analysis shows that coke and 65 fragmented oxygenates and/or olefins. According to Carlson et
char are not graphitic carbon, and should not be treated as such al.,28 although these reactions are promoted by acid catalysts, they
in the modelling of pyrolysis mechanisms.72,73 Further can still occur in the absence of catalyst. Thus, the dominant
10 verification of this observation was performed in TPO mechanisms in catalytic pyrolysis are similar (and co-exist) with
experiments with mixtures of graphitic carbon and coked those of thermal pyrolysis and the mechanism of formation of
catalysts (not shown here), in which the TPO shows one clear 70 coke (defined here as a catalytic product) should be similar to that
DTG peak for the carbon (at ~800°C), well-separated from those of char (defined as a thermal product).
of char and coke.
15
Analysis of the nature of coke deposited on HZSM-5 in pyrolysis
In-depth study of the char/coke composition was also performed of bio-oil (produced from pine) performed by Valle et al.76 and
in Raman. Fig.11 shows the Raman spectra of the char/coke from 75 Ibáñez et al.77 showed that catalyst deactivation is caused by
glucose pyrolysis. The deconvolution of the spectra is based on coke deposition. They identified coke of two origins: thermal and
observation of the main peaks, shoulders, valleys, and tails. Thus catalytic; and observed catalytic coke to be deposited mainly
20 for the slow pyrolysis, the spectra has been deconvoluted into 6 inside the zeolite crystal channels; whereas thermal coke
Gaussian peaks: #1, ca. 1700 cm-1 (D2 band, corresponding to deposited on the outer surface of the catalyst (meso-, macro-
carbonyl groups); #2, ca. 1600 cm-1 (G band, corresponding to 80 pores). This is in agreement with Cheng and Huber,13 who
well-structured aromatic rings); #3, 1550-1570 cm-1 (Gꞌ band, postulated that coke from furan pyrolysis comprises polycyclic
corresponding to coexistence of a well-structured and a not well- oxygenates. The small fraction of soluble coke that they were
25 structured carbons); #4, ca. 1450 cm-1 (D3 band, corresponding to able to extract from the organics retained inside the catalyst had
structural defects of aromatic clusters); #5, 1360-1370 cm-1 (D molecular weights as high as 204 gr/mol, mostly containing
band, corresponding to not well-structured aromatics); and #6, ca. 85 aromatics rings and carbonyl end groups. Pure polyaromatics
1270 cm-1 (corresponding to C-H vibrations). For the fast were also identified. In summary, what we know from the
pyrolysis, 5 peaks were sufficient for fitting the spectra: #1 (D2 relevant literature and this study is that char and coke are not
30 band); #2 (G band); #3 (Gꞌ band); #4 (D band); and #5 (C-H graphite-type carbons. Char shows low TPO temperature peaks,
vibrations). Detailed description of the interpretation of each significant amounts of carbonyl groups and a large number of
band identified by Raman can be found elsewhere.74–76 90 Raman defects. Coke exhibits much higher TPO peaks, shows
Comparison of the Raman spectra of the slow and fast pyrolysis smaller amounts of carbonyl groups and significantly smaller
shows that the second broad peaks (D3+D+C-H in slow pyrolysis; Raman defects. Of course, coke and char co-exist in the solid
35 D+C-H in fast pyrolysis) decrease significantly when fast heating residue of fast catalytic pyrolysis, but as shown in Fig.8, coke is
rates are applied in pyrolysis. The fast catalytic pyrolysis the dominant carbon form after fast catalytic pyrolysis (ratio of
char/coke has the lowest second peak (0.21 in area fraction) 95 7/3). Brewer et al.78 performed 13C-NMR analysis on the chars
among the four. It appears that the formation of defect aromatics obtained from fast and slow thermal pyrolysis of switchgrass,
(mainly D band and D3 band) is not favoured by the fast heating showing that aromatic clusters of 7-8 rings terminated by
40 rates. Combining the results from TPO and Raman implies that carbonyl and hydroxyl groups are the representative composition
the second peak is characteristic of the char formed. This is in of thermal char (a structure similar to the one shown in Fig.12).
agreement with Sheng,75 who claimed that as the defect bands 100 Formation of char is typically observed in concert with CO and
increase, the ordering of the char decreases and the char becomes CO2, which can be explained by the decarbonylation of furanics
more reactive. Moreover, the sum of areas of the G and Gꞌ bands postulated by Huber and co-workers.13,79
45 increases for the fast pyrolysis chars and is highest for the fast

10 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society of Chemistry [year]
View Article Online
Page 11 of 15 Green Chemistry DOI: 10.1039/C3GC41581C

D fSS- a01S  E fSS- a01S  F fSS- a01S  G fSS- a01S 
0SSSS068.o2SSSSSSSSS7o7. 0SSSS068.o2SSSSSSSSS7o70 0SSSS0477o2SSSSSSSSS7o75 0SSSS0683o3SSSSSSSSS7o7.
.SSSS0679o5SSSSSSSSS7o.5 .SSSS0679o5SSSSSSSSS7o53 .SSSS0676o6SSSSSSSSS7o33 .SSSS060.o9SSSSSSSSS7o33
5SSSS0264o2SSSSSSSSS7o.. 5SSSS0264o2SSSSSSSSS7o06 5SSSS0225o8SSSSSSSSS7o09 5SSSS0227o5SSSSSSSSS7o55
3SSSS0333o3SSSSSSSSS7o03 3SSSS0333o3SSSSSSSSS7o03 3SSSS0582o0SSSSSSSSS7o55 3SSSS0596o7SSSSSSSSS7o.7
2SSSS0566o2SSSSSSSSS7o02 2SSSS0566o2SSSSSSSSS7o04 2SSSS0.43o5SSSSSSSSS7o7. 2SSSS0.82o7SSSSSSSSS7o70
6SSSS0.45o7SSSSSSSSS7o.5 6SSSS0.45o7SSSSSSSSS7o09
m 5
 m 
"a#
"a#
. .
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

.777 0977 0677 0377 0.77 0777 .777 0977 0677 0377 0.77 0777 .777 a00977 0677 0377 0.77 0777 .777 0977 0677 0377 0.77 0777
 SS- 1
Fig.11 Raman spectra of glucose char from (a) slow thermal, (b) slow catalytic, (c) fast thermal and (d) fast catalytic pyrolysis. The calculated area
fractions of deconvoluted peaks (#1-6 for slow pyrolysis, #1-5 for fast pyrolysis) in each case are shown.

Green Chemistry Accepted Manuscript


Another interesting observation was reported by Mettler et al.,62 On the basis of the aforementioned possible reaction pathways
5 who compared the char yield from pyrolysis of glucose-based for the formation of coke and char in the catalytic pyrolysis of
carbohydrates and the fraction of carbonyl groups present in the 50 biomass, we suggest representative reaction schemes that are
volatile products and observed a very strong relation. They relevant to the pyrolysis of glucose and cellulose. Four reaction
postulated that aldol condensation chemistry is significant in pathways (noted as RA, RB, RC, and RD) for coke and char
thermal pyrolysis char formation. In parallel to the formation are illustrated in Fig.12, together with the minimum
10 aforementioned studies, there is a wealth of information and projection diameter (MPD) for each component calculated in
published work on the nature and possible mechanisms of coke 55 ChemAxon Marvin.85 Toluene and furfural are chosen as the
formation in the catalytic cracking of gasoil and other starting molecules for showing the mechanisms, since their
hydrocarbons.80–83 Hydrogen transfer and carbenium ion presence as biomass (glucose) pyrolysis products is confirmed in
chemistry is shown to be dominant in the mechanisms of catalytic our liquid product analysis and the study by Carlson et al.28
15 coke formation, having small olefins and single ring aromatics as (1E,3E)-penta-1,3-diene-1,3,5-triol and (2E,4Z)-hepta-2,4-diene-
its origin. The production of aromatics is established in catalytic 60 1,4,7-triol are used as char precursors, as they are (according to
pyrolysis of biomass, while small olefins are mostly postulated as the glucose/cellulose pyrolysis mechanism proposed by Vinu and
responsible for their formation via the hydrocarbon pool Broadbelt73) among the postulated anhydrosugars responsible for
mechanism proposed by Huber and coworkers.28 Therefore, it is carbon formation (char is tracked as carbon in that model). The
20 reasonable to assume that the mechanisms accepted in catalytic reactions of Fig.12 are in no way exhaustive and can be
pyrolysis of hydrocarbons may well be relevant to the catalytic 65 considered only as representative of the mechanisms involved in
pyrolysis of biomass. In that view, there should be a competition coke and char formation. Moreover, there are alternative
between the reactions forming aromatics and the reactions pathways that can co-exist (for instance, dehydration and
responsible for catalytic coke, given the hydrogen-poor tautomerization of the alcohols product in pathways RC and RD),
25 environment in the hydrocarbon pool and the small evidence of which are omitted. The focus of this analysis is to explore the
larger olefins observed in biomass catalytic pyrolysis 70 likely mechanisms that can produce fused or linear polyaromatic
experiments. On the other hand, char formation reactions can be rings with or without oxygen containing ending groups.
surmised to proceed via reactions of oxygenates (aldol and Diels-
Alder), but given the FTIR and Raman evidence they must lead to Mechanisms RA and RB are similar to those proposed in
30 large polyaromatic rings terminated with carbonyl and hydroxyl Cerqueira et al.82 and Quintana-Solórzano et al.,83 respectively,
groups. One other aspect to consider is the steric constraints of 75 and present common coke formation reactions (with hydrocarbon
the ZSM-5 catalyst, which impose a requirement for the catalytic precursors). According to Guisnet & Magnoux,80 the composition
coke to be smaller (in its number of aromatic rings) than the char of high temperature (400-600°C) coke is practically independent
formed. The largest polyaromatic that can be accommodated and of the reactant and is considered to comprise polyaromatics,
35 trapped inside the channel intersections of MFI zeolites is formed not only by condensation and rearrangement steps but
methylpyrene (shown with molecular modelling by Guisnet et 80 also via various hydrogen transfer steps on acid catalysts.
al.84). The formation of polyaromatic rings is accepted to proceed Methylpyrene is a major coke component with a molecular size
via a series of alkylation and hydrogen transfer steps.84 Marin and intermediate between that of the MFI supercages and of its pore
co-workers83 show that coke formation proceeds via hydrogen apertures. Therefore, the intermediate size aromatic rings derived
40 transfer, alkylation and ring closing, with toluene and propylene via pathways RA and RB are sterically blocked inside the
as possible coke precursors. Cheng and Huber,79 accept a 85 supercages of the ZSM-5, while they are thermodynamically
mechanism that produces allene and/or methylacetylene (formed resistant to cracking. Further alkylation and aromatization of the
by decarbonylation of furanics), which undergo oligomerization representative final polyaromatic of the RA mechanism is not
to form a series of olefins and single ring aromatics. Accepting feasible, as it is sterically constrained by the size of the ZSM-5
45 the thermodynamic stability of polyaromatics, it is straight- cavities. First et al.86 report a cavity size of 7.5Å for calcined
forward to envision that hydrogen transfer chemistry is relevant 90 ZSM-5,87 which can tightly accommodate the pyrene molecule
to coke formation in biomass catalytic pyrolysis. (~9Å), given that the ZSM-5 cage and pyrene are not spheres.

This journal is © The Royal Society of Chemistry [year] Journal Name, [year], [vol], 00–00 | 11
View Article Online
Green Chemistry DOI: 10.1039/C3GC41581C
Page 12 of 15

O OH

HO HO HO
O
MPD=6.16
H
C+ H+ KD=5.50
MPD=6.80 HO
H2O HO
KD=5.85 O
RA.1 RB.1 MPD=7.46 MPD=7.08
RC.1

O RD.1
RC2 HO OH OH
O OH HO
HO
MPD=6.84
MPD=7.74 MPD=7.10 OH
MPD=7.64 HO HO
HO
H2O HO HO OH OH
H2 O HO
OH
RA.2 RC.3 OH
OH OH
HO
RB.2
MPD=9.30 O OH
OH RC.4 OH OH
C+ OH HO
O MPD=9.98
MPD=9.68 MPD=9.56 MPD=10.08
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

MPD=7.24
MPD=8.00 H2O
MPD=8.24 RD.2
OH OH
RC.5 R
HO OH OH HO OH
RA.3 HO
OH OH
OH OH
OH OH OH
RB.3 OH

OH OH OH
MPD=8.16 HO
O HO OH OH HO HO HO

MPD=7.94 OH OH OH HO
MPD=8.20 HO HO HO
OH OH OH
RA.4 RC.6 OH OH
O OH

Green Chemistry Accepted Manuscript


RC.7 OH HO HO OH
OH HO HO OH
MPD=12.16 OH
RB.4 RD.3 OH OH HO
HO

OH
OH
OH HO
HO
HO HO OH OH HO HO OH
O MPD=9.34 OH
MPD=9.46 KD=7.24 MPD=8.20 OH OH
OH
OH OH OH
OH OH MPD>9.84 OH

Fig.12 Possible reaction pathways for coke and char formation. (a) RA – toluene self-alkylation via RB.1 (alkylation), RB.2 (dehydrogenative coupling),
82

RB.3 (isomerization), and RB.4 (hydrogen transfer and repetition of RB.1-RB.5); (b) RB83 – coke formation via RA.1 (alkylation on the nucleus with
carbenium ions), RA.2 (side alkylation and isomerization), RA.3 (cyclization), and RA.4 (repetition of RA.2, RA.3); (c) RC – char formation from
5 furfural via RC.1 (Diels-Alder with propylene), and aldol condensations (RC.2-RC.7); RD – Diels Alder cycloadditions of C-5 and C-6 anhydrosugars via
RD.1 (Diels Alder self- or hetero- cycloaddition), RD.2 (Diels Alder cycloaddition) of the products of RD.1 with the original anhydrosugars, and RD.3
(repetition of RD.2, followed by enol-keto tautomerization to produce carbonyl ending groups and condensation to fused polyaromatic rings terminated by
carbonyl and hydroxyl groups). MPD is the minimum projection diameter of each molecule; KD is the kinetic diameter reported by Jae et al.44

Methyl- or ethyl- anthracene (the final product of pathway RB, The aldol is indeed an interesting pathway, because it can lead to
10 ~8Å diameter) cannot escape the ZSM-5 pore (with pore limiting polyaromatics with carbonyl and hydroxyl end groups which are
diameter of 4.5Å and maximum crystallographic diameter of clearly evident in the FTIR and Raman analyses. It can proceed
~6Å). Therefore, polyaromatics once formed inside the zeolite 45 inside the catalyst pore forming oxo-aromatics (Valle et al.76) or
cage are bound to stay in it and undergo further aromatization continue outside of the zeolite, blocking both the micropores and
leading to ring structures with a maximum of ca. 4 rings. Given the macropores of the catalyst (as observed in the BET analyses,
15 their relatively larger residence time inside the zeolite (and the Table 2).
reactor) these polyaromatics might partially condense to heavier
forms, leading to insoluble coke that is difficult to analyse. In 50 Finally, in regards to char formation (besides the aforementioned
summary, we postulate that the dominant coke formation aldol route), we look at the molecular structure of (1E,3E)-penta-
mechanisms of catalytic cracking of hydrocarbons are relevant in 1,3-diene-1,3,5-triol and (2E,4Z)-hepta-2,4-diene-1,4,7-triol,
20 biomass (glucose in this case) pyrolysis as well. The largest coke proposed by Vinu and Broatbelt73 as major char precursors. These
form possible should be defined by the ZSM-5 steric constraints, compounds (and other similar obtained by enol-keto
through a more in-depth analysis is required, such as the work by 55 tautomerizations) can act as dienes and dienophiles, which points
First at al.,86 describing the interactions between the guest and strongly towards a Diels-Alder route for char formation. In Fig.12
host atoms that capture the particular shape of the molecule we enumerate all the possible products of their Diels-Alder
25 subject to its possible rotations. Molecular projection (MPD) and reactions, omitting the possible tautomerization and condensation
kinetic diameters (KD) are of little use in this regard. steps (which would not significantly affect the aromatic structure
60 of these intermediates). The first products of reaction RD.1 are all
Nonetheless, the coke from glucose pyrolysis is quantitatively significantly larger (>9Å MPD) than the ZSM-5 pores or cavities.
much larger than what one would anticipate from hydrocarbons. Hence, their formation can only occur in the catalyst macropores
30 Hence, there must be a significant contribution from oxygenates and on the catalyst surface. In the absence of catalyst they can
in the formation of coke (besides, their obvious contribution to lead to structures similar to those proposed by Brewer et al.78 and
char, discussed later). According to the results obtained in the 65 it is likely that they are promoted at lower temperatures and slow
char/coke analysis of this study and the postulated mechanisms heating rates, as indicated by the larger transmittance intensity of
for coke formation discussed previously, the most important the FTIR carbonyl group in the glucose char from slow thermal
35 mechanisms for coke formation with origins in oxygenated pyrolysis. If we take into account the relatively narrow TPO DTG
compounds are Diels-Alder and aldol reactions. A scheme curves from slow thermal and slow catalytic pyrolysis of glucose
involving these reactions is depicted in pathway RC of Fig.12. 70 and cellulose and the wide defect bands observed in the Raman
Furfural acting as a diene can react with a small olefin (from the spectra of the chars from slow thermal and slow catalytic
hydrocarbon pool) yielding tolualdehydes, which can undergo pyrolysis of glucose, it is reasonable to assume that char is
40 aldol condensation reactions (inside or outside the ZSM-5 pore) composed of a few oxygen-containing polyaromatics, that are
to form polyaromatics of the form proposed by Brewer et al..78 preferably formed at lower temperatures (or their kinetics are

12 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society of Chemistry [year]
View Article Online
Page 13 of 15 Green Chemistry DOI: 10.1039/C3GC41581C

slower). Knowledge of the type (char or coke), location accessibility to the zeolite, and thus a stronger overall catalyst
(macropores or micropores), and amount (weight fraction) of the effect than that of slow catalytic pyrolysis. The wide shapes of
carbonaceous residues during pyrolysis can shed light to the 60 the TPO curves indicate that catalytic coke corresponds to a range
extent and significance of the two domains shown in Fig.12 and of hydrocarbons or oxygenates, formed via different mechanisms.
5 help to extend the excellent work by Vinu and Boradbelt,73 Cho Common catalytic coke formation mechanisms are proposed for
et al.,22,88 Lin et al.,41 Cai et al.,89 Mettler et al.,43,62 and the glucose coke formation (on the basis of FTIR and Raman
Rangarajan et al.,90 to account for thermal and catalytic reactions observations of diminishing carbonyl groups and decreasing
leading to char and coke. 65 defect bands, respectively), which are assumed to proceed in
parallel to aldol and Diels-Alder reactions of oxygen containing
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

Conclusions biomass intermediates. The catalytic coke reactions are of the


same type and antagonistic in nature to the reactions leading to
10 In summary, the properties and the characteristics of the char and
aromatics, resulting in consumption of single-ring aromatics for
coke derived from thermal and catalytic pyrolysis are different. 70 the production of coke inside the zeolite cage.
The formation of char and coke strongly depends on the biomass
source and also on the pyrolysis conditions (e.g., heating rates).

Green Chemistry Accepted Manuscript


Acknowledgements
This study reveals that when glucose and pine are used as
15 biomass feed in slow catalytic pyrolysis, catalyst deactivation due This material is based upon work partly supported by the
to formation of char and the corresponding loss of accessibility UCONN Faculty Large Grant. The authors thank Dr. Roger
(surface coverage and macropore blocking) becomes dominant. Ristau for help with FIB-STEM analysis. The authors also thank
On the other hand, in the cellulose slow catalytic pyrolysis and all 75 W. R. Grace & Co. for supplying the ZSM-5 catalysts used in this
the fast catalytic pyrolysis experiments, the formation of coke is work. GMB thanks Prof. Christodoulos A. Floudas for useful
20 attributed to catalyst micropore blocking. In fast catalytic discussions on the use of ZEOMICS and for modelling the
pyrolysis of glucose and pine, formation of catalytic coke calcined ZSM-5.
proceeds in parallel to the formation of thermal char. We
identified the char to coke ratio to be 3/7 (mass basis) for the case References
of glucose and it is theoretically feasible to do so for other
80 Department of Chemical & Biomolecular Engineering, University of
25 biomass model compounds. The following conclusions were Connecticut, Storrs, 191 Auditorium Road, Unit 3222, Storrs, CT, 06269-
drawn in this work: 3222, USA. Tel: +1-860-486-4602; E-mail: george.bollas@ uconn.edu.
• In slow thermal pyrolysis, TPO results show that, for glucose 1. D. M. Alonso, J. Q. Bond, and J. A. Dumesic, Green Chemistry, 2010,
and pine, similar oxidation reactivities of the char/coke products 12, 1493–1513.
85 2. H. Cui and J. R. Grace, Bioresource technology, 2008, 99, 4008 –
are obtained, compared to the corresponding slow catalytic 4020.
30 char/coke. For cellulose, the oxygen content of the char in slow 3. A. V. Bridgwater and G. V. C. Peacocke, Renewable and Sustainable
thermal pyrolysis is higher than that in slow catalytic pyrolysis. Energy Reviews, 2000, 4, 1–73.
Glucose char is likely produced via thermal Diels-Alder and aldol 4. A. C. Kokossis and A. Yang, Computers & Chemical Engineering,
90 2010, 34, 1397–1405.
reactions, while the origin of cellulose char also includes 5. J. Adam, E. Antonakou, A. Lappas, M. Stöcker, M. H. Nilsen, A.
pyrolytic decomposition. Analysis of the TPO DTG peak shapes Bouzga, J. E. Hustad, and G. Øye, Microporous and Mesoporous
35 shows that cellulose char oxidation has a lower reaction order Materials, 2006, 96, 93–101.
than pine and glucose; indicating the formation of a very narrow 6. G. W. Huber, S. Iborra, and A. Corma, Chemical reviews, 2006, 106,
95 4044–4098.
distribution of hydrocarbons and/or oxygenates. 7. G. W. Huber, National Science Foundation - Division of Chemical,
• In slow catalytic pyrolysis, SEM, STEM and BET results Bioengineering, Environmental, and Transport Systems, 2008.
confirm that macropore blocking occurs for glucose and pine, 8. S. N. Naik, V. V. Goud, P. K. Rout, and A. K. Dalai, Renewable and
40 leading to minor accessibility of the volatiles to the catalyst, Sustainable Energy Reviews, 2010, 14, 578–597.
100 9. D. Mohan, C. U. Pittman,, and P. H. Steele, Energy & Fuels, 2006,
while micropore blocking occurs in the case of cellulose due to 20, 848–889.
coking. Similar reaction orders for the char/coke oxidation are 10. F. A. Agblevor, O. Mante, N. Abdoulmoumine, and R. McClung,
observed in TPO, compared to slow thermal pyrolysis. Energy & Fuels, 2010, 24, 4087–4089.
Application of the random pore model shows that, at slow 11. A. V. Bridgwater, D. Meier, and D. Radlein, Organic Geochemistry,
105 1999, 30, 1479–1493.
45 catalytic pyrolysis of glucose, catalyst contributes only on the 12. A. J. Foster, J. Jae, Y.-T. Cheng, G. W. Huber, and R. F. Lobo,
shape, surface area and pore length of the char formed, but not on Applied Catalysis A: General, 2012, 423-424, 154–161.
its chemical composition. 13. Y.-T. Cheng and G. W. Huber, ACS Catalysis, 2011, 1, 611–628.
• In fast thermal pyrolysis, with TPO DTG peaks shifting to 14. J. Jae, G. A. Tompsett, Y. Lin, T. R. Carlson, J. Shen, T. Zhang, B.
110 Yang, C. E. Wyman, W. C. Conner, and G. W. Huber, Energy &
lower temperatures, glucose and pine char/coke have higher Environmental Science, 2010, 3, 358–365.
50 oxygen content than the corresponding fast catalytic ones. 15. T. R. Carlson, Y.-T. Cheng, J. Jae, and G. W. Huber, Energy &
Comparison of the TPO results of fast thermal pyrolysis char to Environmental Science, 2011, 4, 145.
those from fast catalytic pyrolysis shows that there is no coke 16. A. Atutxa, R. Aguado, A. G. Gayubo, O. Martinand, and J. Bilbao,
115 Energy & Fuels, 2005, 19, 765–774.
formation in fast thermal pyrolysis. Hemicellulose is proposed to 17. M. Guerrero, M. P. Ruiz, M. U. Alzueta, A. R. Bilbao, and A. Millera,
be the most significant char formation precursor in fast thermal Journal of Analytical and Applied Pyrolysis, 2005, 74, 307–314.
55 pyrolysis of pine, on the basis of TPO deconvolution results. 18. J. Cho, J. M. Davis, and G. W. Huber, ChemSusChem, 2010, 3,
• In fast catalytic pyrolysis, as shown by BET, micropore 1162–5.
120 19. G. Elordi, M. Olazar, G. Lopez, P. Castaño, and J. Bilbao, Applied
blocking occurs for all the feedstocks, reflecting good Catalysis B: Environmental, 2011, 102, 224–231.

This journal is © The Royal Society of Chemistry [year] Journal Name, [year], [vol], 00–00 | 13
View Article Online
Green Chemistry Page 14 of 15
DOI: 10.1039/C3GC41581C

20. G. W. Huber and A. Corma, Angewandte Chemie (International ed. 53. E. Apaydın-Varol and A. E. Pütün, Journal of Analytical and Applied
in English), 2007, 46, 7184–201. Pyrolysis, 2012, 98, 29–36.
21. K. S. Triantafyllidis, E. F. Iliopoulou, E. V. Antonakou, A. A. Lappas, 75 54. T. Pattanotai, H. Watanabe, and K. Okazaki, Fuel, 2013, 104, 468–
H. Wang, and T. J. Pinnavaia, Microporous and Mesoporous 475.
5 Materials, 2007, 99, 132–139. 55. A. Zhurinsh, J. Zandersons, and G. Dobele, Journal of Analytical and
22. J. Cho, J. M. Davis, and G. W. Huber, ChemSusChem, 2010, 3, Applied Pyrolysis, 2005, 74, 439–444.
1162–5. 56. H. Zhang, Y.-T. Cheng, T. P. Vispute, R. Xiao, and G. W. Huber,
23. D. J. Nowakowski, a. V. Bridgwater, D. C. Elliott, D. Meier, and P. 80 Energy & Environmental Science, 2011, 4, 2297.
de Wild, Journal of Analytical and Applied Pyrolysis, 2010, 88, 53– 57. Z. Wang, F. Wang, J. Cao, and J. Wang, Fuel Processing Technology,
10 72. 2010, 91, 942–950.
24. M. Al-Haddad, E. Rendek, J.-P. Corriou, and G. Mauviel, Energy & 58. P. R. Patwardhan, J. A. Satrio, R. C. Brown, and B. H. Shanks,
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

Fuels, 2010, 24, 4689–4692. Journal of Analytical and Applied Pyrolysis, 2009, 86, 323–330.
25. M. J. Antal and M. Grønli, Industrial & Engineering Chemistry 85 59. A. Aho, N. Kumar, K. Eränen, T. Salmi, M. Hupa, and D. Y. Murzin,
Research, 2003, 42, 1619–1640. Fuel, 2008, 87, 2493–2501.
15 26. C. Di Blasi, Progress in Energy and Combustion Science, 2008, 34, 60. J. Shen, X.-S. Wang, M. Garcia-Perez, D. Mourant, M. J. Rhodes,
47–90. and C.-Z. Li, Fuel, 2009, 88, 1810–1817.
27. H. Zhang, T. R. Carlson, R. Xiao, and G. W. Huber, Green Chemistry, 61. M. S. Mettler, S. H. Mushrif, A. D. Paulsen, A. D. Javadekar, D. G.
2012, 14, 98. 90 Vlachos, and P. J. Dauenhauer, Energy & Environmental Science,

Green Chemistry Accepted Manuscript


28. T. R. Carlson, J. Jae, Y.-C. Lin, G. A. Tompsett, and G. W. Huber, 2012, 5414.
20 Journal of Catalysis, 2010, 270, 110–124. 62. M. S. Mettler, D. G. Vlachos, and P. J. Dauenhauer, Energy &
29. A. Aho, N. Kumar, K. Eranen, T. Salmi, M. Hupa, and D. Murzin, Environmental Science, 2012, 5, 7797–7809.
Fuel, 2008, 87, 2493–2501. 63. J. Fermoso, M. V. Gil, C. Pevida, J. J. Pis, and F. Rubiera, Chemical
30. M. Asadullah, S. Zhang, Z. Min, P. Yimsiri, and C.-Z. Li, 95 Engineering Journal, 2010, 161, 276–284.
Bioresource technology, 2010, 101, 7935–7943. 64. K. Slopiecka, P. Bartocci, and F. Fantozzi, Applied Energy, 2012, 97,
25 31. M. S. Hasan Khan Tushar, N. Mahinpey, A. Khan, H. Ibrahim, P. 491–497.
Kumar, and R. Idem, Biomass and Bioenergy, 2012, 37, 97–105. 65. J. Adam, M. Blazsό, E. Mésázros, M. Stöcker, M. H. Nilsen, A.
32. C. Di Blasi, Progress in Energy and Combustion Science, 2009, 35, Bouzga, J. E. Hustad, M. Grønli, and G. Øye, Fuel, 2005, 84, 1494–
121–140. 100 1502.
33. D. Ferdous, A. K. Dalai, S. K. Bej, and R. W. Thring, Energy & 66. A. K. Burnham and R. L. Braun, Energy & Fuels, 1999, 13, 1–22.
30 Fuels, 2002, 16, 1405–1412. 67. S. K. Bhatia and D. D. Perlmutter, AIChE Journal, 1980, 26, 379–386.
34. D. J. Nowakowski, A. V. Bridgwater, D. C. Elliott, D. Meier, and P. 68. K. Miura, H. Nakagawa, S. Nakai, and S. Kajitani, Chemical
de Wild, Journal of Analytical and Applied Pyrolysis, 2010, 88, 53– Engineering Science, 2004, 59, 5261–5268.
72. 105 69. V. B. Di Marco and G. G. Bombi, Journal of chromatography. A,
35. R. K. Sharma, J. B. Wooten, V. L. Baliga, X. Lin, W. Geoffrey Chan, 2001, 931, 1–30.
35 and M. R. Hajaligol, Fuel, 2004, 83, 1469–1482. 70. D. Vamvuka, E. Kakaras, E. Kastanaki, and P. Grammelis, Fuel,
36. M. Olazar, R. Aguado, M. J. S. José, and J. Bilbao, Journal of 2003, 82, 1949–1960.
Chemical Technology & Biotechnology, 2001, 76, 469–476. 71. S. Xin, H. Yang, Y. Chen, X. Wang, and H. Chen, Fuel, 2013, 113,
37. R. Aguado, M. Olazar, M. J. S. José, G. Aguirre, and J. Bilbao, 110 266–273.
Industrial & Engineering Chemistry Research, 2000, 39, 1925–1933. 72. Y. Zhang, S. Kajitani, M. Ashizawa, and Y. Oki, Fuel, 2010, 89,
40 38. M. Olazar, R. Aguado, J. Bilbao, and A. Barona, AIChE Journal, 302–309.
2000, 46, 1025–1033. 73. R. Vinu and L. J. Broadbelt, Energy & Environmental Science, 2012,
39. S. Du and G. M. Bollas, Fuel, 2013. 5, 9808.
40. A. Aho, N. Kumar, A. V. Lashkul, K. Eränen, M. Ziolek, P. Decyk, T. 115 74. B. Guichard, M. Roy-Auberger, E. Devers, B. Rebours, A. A.
Salmi, B. Holmbom, M. Hupa, and D. Y. Murzin, Fuel, 2010, 89, Quoineaud, and M. Digne, Applied Catalysis A: General, 2009, 367,
45 1992–2000. 1–8.
41. Y.-C. Lin, J. Cho, G. A. Tompsett, P. R. Westmoreland, and G. W. 75. C. Sheng, Fuel, 2007, 86, 2316–2324.
Huber, The Journal of Physical Chemistry C, 2009, 113, 20097– 76. B. Valle, P. Castaño, M. Olazar, J. Bilbao, and A. G. Gayubo,
20107. 120 Journal of Catalysis, 2012, 285, 304–314.
42. P. R. Patwardhan, J. A. Satrio, R. C. Brown, and B. H. Shanks, 77. M. Ibáñez, B. Valle, J. Bilbao, A. G. Gayubo, and P. Castaño,
50 Bioresource technology, 2010, 101, 4646–4655. Catalysis Today, 2012, 195, 106–113.
43. M. S. Mettler, A. D. Paulsen, D. G. Vlachos, and P. J. Dauenhauer, 78. C. E. Brewer, K. Schmidt-rohr, J. A. Satrio, and R. C. Brown, 2009,
Green Chemistry, 2012, 14, 1284. 28.
44. J. Jae, G. A. Tompsett, A. J. Foster, K. D. Hammond, S. M. Auerbach, 125 79. Y.-T. Cheng and G. W. Huber, Green Chemistry, 2012, 14, 3114.
R. F. Lobo, and G. W. Huber, Journal of Catalysis, 2011, 279, 257– 80. M. Guisnet and P. Magnoux, Applied Catalysis A: General, 2001,
55 268. 212, 83–96.
45. E. M. Sulman, V. V. Alferov, Y. Y. Kosivtsov, A. I. Sidorov, O. S. 81. C. H. Bartholomew, Applied Catalysis A: General, 2001, 212, 17–60.
Misnikov, A. E. Afanasiev, N. Kumar, D. Kubicka, J. Agullo, T. 82. H. S. Cerqueira, G. Caeiro, L. Costa, and F. Ramôa Ribeiro, Journal
Salmi, and D. Y. Murzin, Chemical Engineering Journal, 2007, 134, 130 of Molecular Catalysis A: Chemical, 2008, 292, 1–13.
162–167. 83. R. Quintana-Solórzano, J. W. Thybaut, G. B. Marin, R. Lødeng, and
60 46. T. R. Carlson, G. A. Tompsett, W. C. Conner, and G. W. Huber, A. Holmen, Catalysis Today, 2005, 107-108, 619–629.
Topics in Catalysis, 2009, 52, 241–252. 84. M. Guisnet, L. Costa, and F. Ramôa, Journal of Molecular Catalysis,
47. K. Miura and P. L. Silveston, Energy & Fuels, 1989, 3, 243–249. 2009, 305, 69–83.
48. T. E. Mcgrath, W. G. Chan, and R. Hajaligol, Journal of Analytical 135 85. ChemAxon, Marvin Sketch,
and Applied Pyrolysis, 2003, 66, 51–70. Https://www.chemaxon.com/products/marvin/., 2010.
65 49. P. T. Williams and S. Besler, Renewable Energy, 1996, 7, 233–250. 86. E. L. First, C. E. Gounaris, and C. A. Floudas, Langmuir, 2013, 29,
50. H. Yang, R. Yan, H. Chen, C. Zheng, D. H. Lee, and D. T. Liang, 5599–608.
Energy & Fuels, 2006, 20, 388–393. 87. http://helios.princeton.edu/zeomics/.
51. A. Aho, N. Kumar, K. Eränen, B. Holmbom, M. Hupa, T. Salmi, and 140 88. J. Cho, S. Chu, P. J. Dauenhauer, and G. W. Huber, Green Chemistry,
D. Y. Murzin, International journal of molecular sciences, 2008, 9, 2012, 14, 428.
70 1665–75. 89. J. Cai, W. Wu, R. Liu, and G. W. Huber, Green Chemistry, 2013, 15,
52. A. Zabaniotou, O. Ioannidou, E. Antonakou, and A. Lappas, 1331.
International Journal of Hydrogen Energy, 2008, 33, 2433–2444. 90. S. Rangarajan, A. Bhan, and P. Daoutidis, Industrial & Engineering
145 Chemistry Research, 2010, 49, 10459–10470.

14 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society of Chemistry [year]
Page 15 of 15 Green Chemistry
View Article Online
DOI: 10.1039/C3GC41581C
Published on 13 September 2013. Downloaded by University of Wyoming on 21/09/2013 09:12:56.

Green Chemistry Accepted Manuscript


The origins of char and coke in biomass catalytic pyrolysis are identified and mechanisms for their formation
are proposed

You might also like