You are on page 1of 15

Journal of Analytical and Applied Pyrolysis 92 (2011) 224–232

Contents lists available at ScienceDirect

Journal of Analytical and Applied Pyrolysis


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

Screening acidic zeolites for catalytic fast pyrolysis of biomass and its
components夽
David J. Mihalcik, Charles A. Mullen, Akwasi A. Boateng ∗
Eastern Regional Research Center, Agricultural Research Service, U. S. Department of Agriculture, 600 E. Mermaid Lane, Wyndmoor, PA 19038, United States

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

Article history: Zeolites have been shown to effectively promote cracking reactions during pyrolysis resulting in
Received 6 April 2011 highly deoxygenated and hydrocarbon-rich compounds and stable pyrolysis oil product. Py/GC–MS was
Accepted 6 June 2011 employed to study the catalytic fast pyrolysis of lignocellulosic biomass samples comprising oak, corn
Available online 14 June 2011
cob, corn stover, and switchgrass, as well as the fractional components of biomass, i.e., cellulose, hemicel-
lulose, and lignin. Quantitative values of condensable vapors and relative compositions of the pyrolytic
Keywords:
products including non-condensable gases (NCG’s) and solid residues are presented to show how reaction
Catalytic fast pyrolysis
products are affected by catalyst choice. While all catalysts decreased the oxygen-containing products
Zeolite
Pyrolysis oil
in the condensable vapors, H-ZSM-5 was most effective at producing aromatic hydrocarbons from the
pyrolytic vapors. We demonstrated how the Si/Al ratio of the catalysts plays a role in the deoxygenation
of the vapors towards the pathway to aromatic hydrocarbons.
Published by Elsevier B.V.

1. Introduction ratio [33]. Zeolites with higher acidity, or lower Si/Al ratios, have
been shown to be effective in promoting cracking reactions dur-
Owing to its vast availability and potential application as a ing the incipient pyrolysis reactions [34–37]. However, there are as
liquid fuel, the use of biomass for renewable energy production many catalysts as there are types of biomass available so screen-
has received increasingly more attention over the last 40 years[1]. ing of various catalysts/biomass combinations is useful. Py/GC–MS
Among the various conversion technologies being studied, fast serves as a useful tool to understanding pathways leading to sta-
pyrolysis has received significant attention because it is feedstock ble pyrolysis liquids so promising catalysts/biomass candidates can
neutral [2] and has the potential to be easily incorporated into be down-selected for larger experiments. In this study, a series of
a distributed process model [3]. This notwithstanding the major zeolitic catalysts, chosen based on their acidity and framework,
drawback towards commercialization of fast pyrolysis oil (bio-oil) and several types of biomass, including their pure components,
is its instability caused by high oxygen content and acidity, thereby were screened using Py/GC–MS to determine critical degradation
rendering it unsuitable for incorporation into existing petroleum- pathways of in situ catalytic pyrolysis.
based infrastructure “as is” [2,4]. Upgrading technologies that favor The objective of the study was to use Py/GC–MS to screen ligno-
pathways towards reducing the oxygen content are therefore nec- cellulosic and component biomass samples against different types
essary [1]. Among these technologies, incorporation of catalysts of commercially available zeolite catalysts to determine relative
into the pyrolysis reaction, in situ, to reduce the more reactive abilities to deoxygenate pyrolytic condensable components and
oxygenated compounds appears to be most pragmatic [5–8]. Of produce aromatic hydrocarbons.
the various catalysts studied to date, [9–19] zeolites have received
much attention due to their vast availability, relatively low cost, and
2. Materials and methods
facile tunability [20–32]. Zeolite catalysts have been shown to be
effective in the selective deoxygenation of pyrolytic vapors, result-
2.1. Pyrolysis GC-MS
ing in the formation of aromatics and effectively increasing the C/O
Pyrolysis experiments were conducted using a Frontier Lab
Double-Shot micro pyrolyzer PY-2020iD equipped with the
夽 Mention of trade names or commercial products in this publication is solely for Frontier Lab Auto-Shot Sampler AS-1020E coupled to a gas chro-
the purpose of providing specific information and does not imply recommendation matograph, Shimadzu GC-2010. Pyrolysis products were detected
or endorsement by the U.S. Department of Agriculture. USDA is an equal opportunity
provider and employer.
using a Shimadzu GCMS-QP2010S mass spectrometer (MS). The
∗ Corresponding author. Tel.: +1 215 233 6493; fax: +1 215 233 6406. micro pyrolyzer comprised an interface heater (100–400 ◦ C) and
E-mail address: akwasi.boateng@ars.usda.gov (A.A. Boateng). a pyrolysis furnace (40–800 ◦ C). In this setup, approximately 1 mg

0165-2370/$ – see front matter. Published by Elsevier B.V.


doi:10.1016/j.jaap.2011.06.001
D.J. Mihalcik et al. / Journal of Analytical and Applied Pyrolysis 92 (2011) 224–232 225

of biomass is placed at the bottom of a deactivated stainless steel O


O OH HO
cup, followed by the addition of approximately 5 mg of catalyst O O O O O
(∼1:5/w:w, biomass to catalyst ratio). Among the limitations of H HO
OH OH OH
this small scale system is that there is no actual determination
1 2 3 4 5
of the deactivation of the catalyst, which necessitates the large
excess of catalyst. The biomass/catalyst sample is not initially Acetic Acid Furfural Furfural Alcohol Acetol Levoglucosan
mixed and is placed in the Auto-Shot Sampler (48 sample capac- OCH3
ity). The sample is gravity fed into the inert atmosphere of the OH OH OH
pre-heated pyrolysis furnace (550 ◦ C), where the sample is sub-
jected to pyrolysis conditions for 18 s. The helium carrier gas is OCH3 OCH3
also used to purge air in the sample prior to pyrolysis and to con- 6 7 8 9 10
vey the pyrolysis gas through the pyrolysis reactor, a quartz tube, Guaiacol Syringol Phenol Benzene Toluene
then to the GC–MS. The fraction of the pyrolysis vapors that are
not detected by GC can be analyzed by LC–MS, NMR, etc., when
produced on larger scale systems where all condensation products
are collected. To analyze the high molecular weight compounds
11 12 13 14 15
eluted in the pyrolysis vapor, i.e., pyrolysis liquid products that
Ethyl Benzene p-Xylene o-Xylene Naphthalene 1-Methyl
could be collected by condensation, GC analyses were performed Naphthalene
on a DB-1701 60 m × 0.25 mm, 0.25 ␮m film thickness. The oven
was programmed to hold at 45 ◦ C for 4 min and ramped at 3 ◦ C/min Fig. 1. Typical oxygenated compounds associated with fast pyrolysis of biomass
to 280 ◦ C after which it was held at this temperature for 20 min. (1–8). Products of catalytic fast pyrolysis using zeolites are typically aromatic hydro-
The injector temperature was kept at 250 ◦ C with the injector split carbons (9–15).

ratio set to 30:1 and helium flow rate was maintained at 1 mL/min.
A separate experiment was conducted using the same procedure A series of 15 condensable pyrolysis products were grouped
as above for analysis of non-condensable pyrolysis gases but using under oxygenated and aromatic compounds and were evaluated
a 100:1 split ratio. In this case, GC analysis was performed using by analysis of variance to determine the de-oxygenation effects
a fused silica capillary column, CP-PoraBOND Q, 25 m × 0.25 mm of catalyst and feedstock interactions, as well as catalyst abil-
(Varian, Palo Alto, CA) with the following program: 3 min at 35 ◦ C ity to produce aromatic hydrocarbons. All the response variables
then ramped at 5 ◦ C/min up to 150 ◦ C followed by 10 ◦ C/min up were considered with all “0” values replaced by a random number
to 250 ◦ C and held for 45 min for a total run time of 81 min. The between 0 and 0.03 (0.03 is limit of detection) where the effects
permanent gas yields were quantified with calibration curves pro- and interactions deemed to be significant (p < 0.05) were further
duced using a standard gas mixture comprising CO, CO2 , CH4 , C2 H4 , examined by performing mean separations using the Bonferroni
C2 H6 , C3 H8 , and C4 H10 in helium (custom-mixed by Scott Spe- LSD technique. Use of 0.03 is an attempt to keep the large num-
cialty Gases, Plumsteadville, PA). Hydrogen was not included in ber of “0” responses from biasing the estimate of variation. Each
the NCG analysis because it was not able to be detected using response variable has the mean separations presented in two forms
the GC–MS setup, and thus, the total product distribution of the of data comparison. The first type allows for the direct compari-
NCG’s is not presented. Each run was performed in triplicate and son of yield observed for a specific pyrolytic product from a single
relevant compounds were quantitatively analyzed. To quantify the catalyst across all feedstocks. The second type facilitates the com-
solid residue remaining after pyrolysis we carried out manual inser- parison of the effect of all catalysts on the wt% production of single
tions of stainless steel cups containing biomass/catalyst mixture in pyrolytic product for each individual feedstock.
a similar 1:5/w:w, ratio, into the pyrolyzer under a flow of helium
and subjected it to temperatures of 550 ◦ C for 18 s and subse- 3. Results
quent weight loss analysis was performed on an analytical scale.
The yield wt%’s for a total of 15 condensable pyrolysis products, 3.1. Non-catalytic pyrolysis
7 non-condensable gases, and residual solids, were quantified for
each feedstock/catalyst combination. The wt% of the majority of the Fast pyrolysis initiates decomposition of the biomass to pro-
condensable pyrolysis products was then calculated by difference duce a complex mixture of products. The product mixture can
to allow for total % composition to be determined for comparison vary depending on the amounts of each component found
purposes. in the feedstock [39,40]. The efficiency of the catalysts was
monitored by restricting the quantitative analysis to fifteen com-
pounds commonly associated with the catalytic/non-catalytic
2.2. Biomass/catalyst combinations pyrolysis of biomass, eight oxygenated compounds and seven
aromatic hydrocarbons (Fig. 1). Also, several non-condensable
The biomass samples included woody biomass, energy crops, gases (NCGs), with the exception of hydrogen, and condens-
and agricultural residues. Additionally, primary biomass compo- able components of the pyrolysis reactions were measured using
nents, i.e., cellulose (Sigma–Aldrich), hemicellulose (xylan from Py/GC–MS.
beechwood, Sigma–Aldrich), and lignin (“Asian”, from Granit Table S1 presents the statistical analysis and yields of all feed-
Research and Development SA) and a 50/50 mixture of cellulose stocks under non-catalytic pyrolysis conditions. They are consistent
and lignin (ETEK lignin) were screened (Table 1). with previously reported data and will not be discussed in depth
Catalyst used included the activated forms of several commer- here, rather, Table S1 is provided as a baseline for comparison with
cially available zeolites including: H-Mordenite, H-ZSM-5, H-Y, catalytic pyrolysis screening. Consistent with previously reported
H-Beta, and H-Ferrierite, in a five factorial experimental design. results, [41] ∼61.50 wt% levoglucosan was produced following non-
The selected catalysts were obtained from Zeolyst International and catalytic pyrolysis of cellulose. Where the feedstock comprised an
encompassed different frameworks (Table 2) and varying acidities equal mixture of cellulose and lignin in 50:50 ratio (ETEK lignin) the
with Si/Al ratio ranging from 5.1 to 360. All catalysts were activated levoglucosan yield reduced to 12.61 wt% similar to that previously
to their protonated form in a muffle furnace at 400 ◦ C for 5 h. reported by Mullen and Boateng [42]. The non-catalytic pyroly-
226 D.J. Mihalcik et al. / Journal of Analytical and Applied Pyrolysis 92 (2011) 224–232

Table 1
Composition assumed for selected lignocellulosic feedstock materials [38] and approximate amounts of cellulose, hemi-cellulose, and lignin associated with tested feedstocks
and feedstock components.

Feedstock Cellulose fraction Hemi-cellulose fraction Lignin fraction

Energy crops 36.6 16.1 21.9


Crop residues 38.0 32.0 17.0
Woody biomass 43.7 28.3 24.3

Cellulose 100 0 0
Hemi-cellulose (xylan) 0 100 0
Pure lignin (Asian) 0 0 100
ETEK lignin 50 0 50
Oak 47 20 27
Corn cob 30 38 3
Corn stover 48 29 6
Switchgrass 32 28 16

sis products of hemicellulose included increased levels of acetol 3.2.2. H-Y


at 7.06 wt%, the highest amount observed for all feedstocks tested. H-Y zeolite is described as having a cubic structure with a
The non-catalytic pyrolysis of pure lignin, Asian lignin in this case, pore system composed of 12-membered rings. Of all the zeolite
yielded relatively large amounts of guaiacol (0.49 wt%) and syringol catalysts tested, H-Y has the highest acidity (Si/Al = 5.1), largest
(1.01 wt%). Additionally, all feedstock components produced trace surface area (925 m2 /g), and largest average pore size (0.74 nm).
amounts of aromatic hydrocarbons. The quantitative and statistical analyses of products formed over
Non-catalytic pyrolysis of lignocellulosic biomass samples pro- H-Y (5.1) zeolite are presented in Table S5. Upon pyrolysis of all
duced about 6–8 wt% of acetic acid which is significantly more than feedstocks over the catalyst, furfural alcohol and levoglucosan were
is produced by biomass components, 1.8–3.1 wt%. Corn cob had produced in similar amounts and levels of phenol were similar for
increased levels of phenol (0.25 wt%), similar to levels produced lignocellulosic biomass. Otherwise, few statistically relevant trends
by pure lignin (0.20 wt%). Consistent with biomass components, were observed in terms of oxygenate production. Statistically, no
low values were observed for all aromatic hydrocarbons following significant differences in aromatic hydrocarbon compounds were
non-catalytic pyrolysis of lignocellulosic biomass. encountered among all feedstocks, primary or lignocellulosic.
Compared with non-catalytic pyrolysis, the H-Y (5.1) catalyst
influenced pyrolysis of biomass and its primary products. It effec-
3.2. Catalytic pyrolysis over zeolites
tively reduced the production of acetic acid from cellulose, 1.89
to 0.015 wt%, and decreased acetic acid formation from all ligno-
3.2.1. H-Ferrierite and H-Mordenite
cellulosic feedstocks. The presence of acetic acid was significantly
H-Ferrierite and H-Mordenite zeolites have similar acidities
reduced for corn cob from around 8.19 to 3.23 wt%. Likewise, ace-
(Si/Al ratio = 20), surface areas and pore sizes with each having an
tol levels decreased from 5.83 wt% non-catalytically for corn cob
orthorhombic structure and respective 8–10, and 12-membered
to 1.33 wt% with H-Y (5.1). The catalyst was also effective at the
rings. The products of pyrolysis of the cellulosic biomass and
reduction of levoglucosan. Again, the increased observation of sub-
biomass primary compounds over these zeolite catalysts are pre-
stituted furans, suggests the switch from a depolymerization to a
sented in Tables S2 and S3 to discern the effect of catalyst for each
fragmentation mechanism in the presence of H-Y (5.1) catalyst.
feedstock. For both component and lignocellulosic biomass, the use
There was not a substantial increase in the amounts of aromatic
of H-Ferrierite (20) gave similar yields for most oxygenates but
hydrocarbons produced, roughly < 2 wt% increase, a value similar
there were statistical differences in the yields of furfural, guaiacol
to that of the H-Ferrierite (20) catalyst.
and syringol. Oxygenate yields varied less when H-Mordenite (20)
was employed, but statistical differences were more prevalent in
the case of furfural, guaiacol, syringol, and phenol. Very little vari- 3.2.3. H-ZSM-5
ation was observed for either catalyst in terms of the quantified The H-ZSM-5 zeolite consists of a MFI orthorhombic crystal
aromatic hydrocarbon products among both primary compounds structure and is characterized as having a 3D network of intercon-
and the biomass used as feedstocks, except to say that they were nected pores arranged in straight 10-membered ring pore system
different in comparison with non-catalytic pyrolysis. 5.2 × 5.7 Å channels connected by sinusoidal 5.3 × 5.6 Å channels.
A rather significant reduction in acetic acid was observed fol- A Si/Al ratio of 23, combined with a potentially ideal pore size and
lowing the pyrolysis of lignocellulosic biomass over these catalysts, structure, leads to an increased production of aromatic hydrocar-
compared to the component feedstocks as is shown in Table S4. bons from each feedstock and feedstock components tested. The
Compared to non-catalytic pyrolysis, increase in the production of quantitative and statistical analyses of products formed over H-
aromatic hydrocarbons was detected with H-Mordenite (20), more ZSM-5 (23) zeolite are presented in Table 3. Compared to previous
so than with H-Ferrierite (20), across all feedstocks. However, the catalysts described, H-ZSM-5 (23) reduced nearly all oxygenated
increased detection of aromatic hydrocarbons in either case was compounds to statistically similar levels regardless of feedstock.
limited to amounts under 2 wt%. Conversely, there were larger statistical differences for the produc-

Table 2
Characteristics of zeolites tested.

Catalyst (product #) SiO2 /Al2 O3 ratio Pore dimensions (Å) Average pore size (nm) Description

Ferrierite (cp914c) 20 4.2 × 5.4, 3.5 × 4.8 0.39–0.51 8,10-Membered rings (FER) Orthorhombic
Mordenite (cbv21a) 20 6.5 × 7.0, 3.4 × 4.8, 2.6 × 5.7 0.42–0.58 12-Membered rings (MOR) Orthorhombic
Y (cbv300) 5.1 7.4 × 7.4 0.74 12-Membered rings (FAU) Cubic
ZSM-5 (cbv2314, cbv5524, cbv28014) 23, 50, 280 5.1 × 5.5, 5.3 × 5.6 0.52–0.55 10-Membered rings (MFI) Orthorhombic
Beta (cp814e, cp814c, cp811c) 25, 38, 360 6.6 × 6.7 0.61–0.62 10, 12-Membered rings (BEA) Tetragonal
D.J. Mihalcik et al. / Journal of Analytical and Applied Pyrolysis 92 (2011) 224–232 227

Table 3
Quantitative and statistical analysis of the wt% of pyrolytic products (1–15) observed following catalytic pyrolysis of component and lignocellulosic feedstocks over H-ZSM-5
(23); relative values of NCG’s, solid residue, and total calculated condensable vapors using H-ZSM-5 series.a

Cellulose Hemi-cellulose Lignin ETEK Lignin Oak Corn cob Corn stover Switchgrass

Acetic acid 2.25A 2.02A 0.0091B 1.72A 2.31A 3.07A 1.72A 2.17A
Furfural 0.76A 0.0079B 0.022B 0.011B 0.016B 0.013B 0.014B 0.020B
Furfural Alcohol 0.019A 0.013A 0.018A 0.010A 0.023A 0.021A 0.017A 0.010A
Acetol 0.12A 0.12A 0.010A 0.12A 0.93A 0.31A 0.013A 0.020A
Levoglucosan 0.31A 0.022A 0.019A 0.019A 0.48A 0.0094A 0.025A 0.019A
Guaiacol 0.019B 0.010B 0.12AB 0.15A 0.099AB 0.10AB 0.084AB 0.093AB
Syringol 0.013C 0.023C 0.22A 0.018C 0.12B 0.048C 0.037C 0.016C
Phenol 0.14ABC 0.10C 0.15ABC 0.22A 0.16ABC 0.19AB 0.17ABC 0.11BC

Benzene 3.06A 1.48C 1.39C 2.96A 2.93A 3.12A 3.17A 1.86B


Toluene 3.79A 2.99BC 2.68CD 2.24D 2.28D 2.23D 2.17D 3.36AB
Ethyl benzene 1.77B 0.26CD 0.21D 0.30CD 0.37CD 0.41C 0.37CD 2.10A
p-xylene 2.15E 2.59D 2.46DE 3.17C 4.12A 3.49BC 3.65B 2.56D
o-xylene 0.54CD 0.59BC 0.40D 0.63ABC 0.74AB 0.77A 0.77A 0.55CD
Naphthalene 1.44A 0.50C 0.75B 1.33A 1.37A 1.33A 1.33A 0.89B
Methyl Naphthalene 0.85C 0.60D 0.70CD 1.19AB 1.28A 1.16AB 1.12B 0.81C

H-ZSM-5 (23)
NCG’s 29.4 20.4 15.5 26.9 26.5 28.1 40.4 28.9
Solid residue 8.8 12.2 19.8 11.9 13.9 23.4 28.2 19.4
Condensables 61.8 67.4 64.7 61.2 59.6 48.5 31.4 51.7

H-ZSM-5 (50)
NCG’s 32.3 24.7 20.9 23.9 34.9 33.3 31.7 31.1
Solid residue 3.6 18.7 11.4 29.0 20.5 15.3 27.7 15.9
Condensables 64.1 56.6 67.7 47.1 44.6 51.4 40.6 53.0

H-ZSM-5 (280)
NCG’s 16.6 17.1 14.7 21.9 21.3 25.3 26.3 28.1
Solid residue 4.3 20.9 33.2 10.7 8.9 14.0 22.8 22.5
Condensables 79.1 62.0 52.1 67.4 69.8 60.7 50.9 49.4
a
Any two means in the same row with no letter in common are significantly different (p < 0.05) by the Bonferroni least significant difference (LSD) technique. The letters
A–F refer to the highest estimates to the least.

tion of aromatic hydrocarbon compounds, more so from feedstock observed for the less acidic catalysts (Fig. 3) with p-xylene being
components than for lignocellulosic biomass samples. produced in the largest amounts in nearly every case. This trend
As mentioned above, H-ZSM-5 (23) reduced the production of indicates that the relative acidities of the zeolites play a larger role
oxygenates to similar levels, regardless of feedstock. Like other in the formation of aromatics than in the processing of oxygenates.
catalysts, H-ZSM-5 (23) drastically reduced the amount of levoglu-
cosan found in the pyrolysis products of cellulose to levels below
3.2.4. H-Beta
0.31 wt% compared to non-catalytic results (Table S4). Although,
H-Beta zeolite has an acidity number of (25) and is composed
substituted furans were not observed as with previous catalysts
of a tetragonal crystal structure with straight 12-membered ring
most likely due to the increased cracking ability of H-ZSM-5 (23).
channels (7.6 × 6.4 Å) with crossed 10-membered ring channels
Upon pyrolysis of hemicellulose over H-ZSM-5 (23) levels of ace-
(5.5 × 6.5 Å). The quantitative and statistical analyses of products
tol dropped from 7.06 wt% for non-catalytic pyrolysis, to 0.12 wt%.
formed over H-Beta (25) zeolite are presented in Table 4. Statis-
Pyrolysis of lignin over H-ZSM-5 (23) resulted in reduction of acetic
tically, H-Beta (25) was equally effective at reducing oxygenated
acid to trace amounts while syringol was reduced from 1.01 to
compounds from pyrolysis of component and lignocellulosic feed-
0.22 wt%.
stocks. Statistically, larger differences were observed for the
H-ZSM-5 (23) increased the production of aromatic hydrocar-
production of aromatic hydrocarbons from biomass components
bons in every case when compared to non-catalytic pyrolysis and
when compared to lignocellulosic biomass samples. Yields of aro-
was the most successful of any catalyst tested (Table S4). The
matic hydrocarbons were significantly larger than all catalysts
presence of H-ZSM-5 (23) also increased the likelihood of the for-
except for H-ZSM-5 (23).
mation of substituted benzenes and naphthalenes, especially when
pyrolyzing cellulose. Furthermore, when pyrolyzing component
feedstocks over H-ZSM-5 (23), toluene was produced in the great- H-ZSM-5 (23) H-ZSM-5 (50) H-ZSM-5 (280)
est abundance and is highest for cellulose at 3.79 wt%. For H-ZSM-5 10
9
(23) over lignocellulosic biomass, with the exception of switch-
wt % Feedstock

8
7
grass, p-xylene was the most abundant hydrocarbon produced, and 6
was highest for oak (4.12 wt%). 5
4
Because H-ZSM-5 (23) was effective at reducing the oxygen con- 3
2
tent of pyrolytic vapors and increasing the production of aromatic 1
0
hydrocarbons, a series of H-ZSM-5 catalysts with Si/Al ratios of 50
and 280, were compared. As Fig. 2 shows, with the exception of
corn stover, the abundance of oxygenates was lowest when using
the H-ZSM-5 (50) catalyst, while the presence of oxygenates was
highest with H-ZSM-5 (280). On the other hand, total aromatics for-
Fig. 2. Comparison of the wt% of oxygenated compounds 1–8 produced by H-ZSM-5
mation from H-ZSM-5 (23) was more than three times the amount zeolites.
228 D.J. Mihalcik et al. / Journal of Analytical and Applied Pyrolysis 92 (2011) 224–232

Table 4
Quantitative and statistical analysis of the wt% of pyrolytic products (1–15) observed following catalytic pyrolysis of component and lignocellulosic feedstocks over H-Beta
(25); relative values of NCG’s, solid residue, and total calculated condensable vapors using H-Beta series.a

Cellulose Hemi-Cellulose Lignin ETEK lignin Oak Corn cob Corn stover Switchgrass

Acetic acid 2.01A 1.92A 0.017B 1.75A 0.0082B 1.76A 1.90A 2.23A
Furfural 0.68A 0.018B 0.018B 0.016B 0.025B 0.016B 0.013B 0.23B
Furfural Alcohol 0.020A 0.0080A 0.021A 0.018A 0.017A 0.021A 0.012A 0.016A
Acetol 0.35A 0.58A 0.0046A 0.14A 0.0088A 0.12A 0.016A 0.11A
Levoglucosan 0.013A 0.012A 0.016A 0.018A 0.14A 0.0093A 0.018A 0.021A
Guaiacol 0.018A 0.021A 0.13A 0.11A 0.019A 0.080A 0.077A 0.073A
Syringol 0.019A 0.0087A 0.063A 0.023A 0.014A 0.010A 0.016A 0.016A
Phenol 0.10A 0.083A 0.16A 0.12A 0.087A 0.11A 0.11A 0.10A

Benzene 1.12BCD 0.84D 1.33ABC 1.48AB 1.44AB 1.51A 1.43AB 0.96CD


Toluene 1.47BC 1.27C 1.83AB 1.95AB 2.05A 2.17A 2.21A 1.44BC
Ethyl benzene 0.62A 0.62A 0.17B 0.20B 0.20B 0.25B 0.26B 0.66A
p-xylene 0.73C 0.73C 0.96ABC 0.97ABC 1.10A 1.09AB 1.21A 0.78BC
o-xylene 0.32A 0.31A 0.39A 0.38A 0.43A 0.43A 0.47A 0.33A
Naphthalene 0.40CD 0.38D 0.70A 0.53BCD 0.60AB 0.63AB 0.56ABC 0.48BCD
Methyl Naphthalene 0.18C 0.25BC 0.43A 0.30ABC 0.35AB 0.39AB 0.37AB 0.15C

H-Beta (25)
NCG’s 30.9 24.8 18.2 26.2 28.4 31.9 30.3 27.8
Solid residue 5.4 22.4 42.9 23.1 29.5 20.9 23.1 28.8
Condensables 63.7 52.8 38.9 50.7 42.1 47.2 46.6 43.4

H-Beta (38)
NCG’s 31.9 23.8 18.6 24.1 30.7 33.3 26.1 27.8
Solid residue 11.6 12.4 44.6 24.7 40.1 34.3 32.9 29.0
Condensables 56.5 63.8 36.8 51.2 29.2 32.4 41.0 43.2

H-Beta (360)
NCG’s 20.4 20.1 18.9 23.9 27.3 27.7 25.5 24.9
Solid residue 7.2 12.8 11.6 18.4 18.5 27.4 17.5 26.5
Condensables 72.4 67.1 69.5 57.7 54.2 44.9 57.0 48.6
a
Any two means in the same row with no letter in common are significantly different (p < 0.05) by the Bonferroni least significant difference (LSD) technique. The letters
A–F refer to the highest estimates to the least.

For all feedstocks, levoglucosan levels were brought below H-Beta (25) H-Beta (38) H-Beta (360)
4
0.026 wt%. The lignocellulosic biomass samples showed a marked 3.5
decrease in the levels of acetic acid and furfural compared to non- 3
wt % Feedstock

catalytic pyrolysis. As is the case with H-ZSM-5 (23), there was 2.5
2
an increased presence of substituted benzenes and naphthalenes,
1.5
while aromatic hydrocarbon production increased in all instances 1
with biomass components but the amount produced were less than 0.5
those achieved with H-ZSM-5 (23). Toluene was the most abundant 0

hydrocarbon produced for all lignocellulosic feedstocks as well as


lignin and ETEK lignin (Table S4).
In Fig. 4, the reduction in oxygenates was investigated using
an H-Beta zeolite series with Si/Al = 25, 38, 360. The acidity differ- Fig. 4. Comparison of the wt% of oxygenated compounds 1–8 produced by H-Beta
ence has little effect on ability to reduce oxygenates with regard to zeolites.
cellulose and hemicellulose feedstocks. However, a more dramatic
effect between H-Beta catalysts was observed for lignin, oak, corn
cob, and corn stover. As shown in Fig. 5, H-Beta (25) was effective at
producing aromatics at levels of 4.5–6.5 wt% for all feedstocks. Use from 1.5 to 2.5 wt%. Lastly, the use of H-Beta (360) was limited in
of H-Beta (38) decreased aromatics production to levels ranging its ability to produce aromatics where levels were around 1.0 wt%.

H-ZSM-5 (23) H-ZSM-5 (50) H-ZSM-5 (280) H-Beta (25) H-Beta (38) H-Beta (360)
18 8
16
wt % Feedstock

7
14
6
wt % Feedstock

12
10 5
8 4
6 3
4
2 2
0 1
0

Fig. 3. Comparison of the wt% of aromatic hydrocarbons 9–15 produced by H-ZSM-5 Fig. 5. Comparison of the wt% of aromatic hydrocarbons 9–15 produced by H-Beta
zeolites. zeolites.
D.J. Mihalcik et al. / Journal of Analytical and Applied Pyrolysis 92 (2011) 224–232 229

100% 100%
90% 90%
80%
80%
% Composion

70%
70%

% Composion
60%
NCGs
50% 60%
Aromacs 9-15 NCGs
40% 50%
Other Condensibles Aromacs 9-15
30% 40%
20% Solids Other Condensables
30%
10% Solid
20%
0%
No Catalyst H-Ferrierite H-Mordenite H-ZSM-5 (23) H-Beta (25)
10%
(20) (20) 0%
No H-ZSM-5 H-ZSM-5 H-ZSM-5
Fig. 6. Average total compositional wt% of pyrolytic condensables, NCG’s, and solid Catalyst (280) (50) (23)
residue over lignocellulosic feedstocks for each catalyst.
Fig. 7. Average total compositional wt% of pyrolytic condensables, NCG’s, and solid
residue over lignocellulosic feedstocks when using H-ZSM-5 catalysts.

3.3. Total compositional analysis of catalytic pyrolysis products


100%
3.3.1. Comparison of zeolites 90%
Fig. 6 depicts the average compositional analyses for lignocellu- 80%
losic biomass feedstocks. Non-catalytic pyrolysis produces about 70%

% Composion
70 wt% condensables on average for all feedstocks tested. This 60%
NCGs
value is less for each of the catalysts and smallest for H-Beta 50%
Aromacs 9-15
(25) (∼48 wt%). Gas production associated with catalytic and non- 40%
Other Condensables
catalytic pyrolysis was characterized by CO, CH4 , CO2 , C2 H4 , C2 H6 , 30%
20% Solid
C3 H6 , C3 H8 , and C4 H10 . A distinct increase in non-condensable gas
production is observed for all feedstocks over all catalysts when 10%

compared to non-catalytic pyrolysis which averaged 14 wt%. NCG 0%

ranged from a low of 19.9 wt% for H-Mordenite (20) to a high of No H-Beta H-Beta H-Beta
Catalyst (360) (38) (25)
30.9 wt% for H-ZSM-5 (23). These trends are also consistent with
catalytic pyrolysis over zeolite where bio-oil yields are consider- Fig. 8. Average total compositional wt% of pyrolytic condensables, NCG’s, and solid
ably reduced and the conversion of the oxygenated species in the residue over lignocellulosic feedstocks when using H-Beta catalysts.
oils was largely due to H2 O loss at lower catalyst temperatures and
CO and CO2 formation at higher catalyst temperatures [5]. ties of the catalysts or on the amounts of aromatic hydrocarbons
Correspondingly, use of H-Beta yielded the larger amount of produced during the 18 s pyrolysis time period.
coke (25.5 wt%), which most likely contributed to premature deac-
tivation of the catalyst. H-ZSM-5 (23) produced the least residue
3.3.3. Comparison of H-Beta series
of all catalysts, about 21.2 wt%, while non-catalytic reactions pro-
Fig. 8 compares the average compositional analysis of the H-Beta
duced 15.8 wt% of feedstock on average. This indicates that even
catalyst series compared to non-catalytic runs. The average solid
with the large amount of aromatic hydrocarbons produced by
production for the series in the order of increasing acidity is 22.4,
H-ZSM-5 (23) (12.6 wt%), instantaneous coking was kept to a min-
34.0, and 25.5 wt% of biomass sample. The large amount of solids
imum. H-Beta (25) causes a large amount of coking of the pyrolysis
formed with H-Beta (38) leads to the least amount of condensables
vapors. On the other hand, H-Ferrierite (20) and H-Mordenite (20)
being formed, 36.4 wt%, of all Beta catalysts. The amount of solid
lacked reactivity towards the pyrolytic vapors presumably due to
residues recovered increased in all cases and was most significant
framework characteristics associated with the pore system of the
for oak where solid amounts increased from 16.2 to 29.5 wt%. The
catalysts.
relatively small increase in solid residue when comparing H-Beta
(360) to non-catalytic values may be attributed to the catalysts rel-
3.3.2. Comparison of H-ZSM-5 series ative inactivity. NCG formation tends to increase with increased
Fig. 7 compares the average compositional analysis of H-ZSM-5 catalytic activity, or increasing acidity within the series.
catalysts and their effects on the pyrolytic vapors of lignocellu-
losic feedstocks. The total amount of condensables is estimated as 4. Discussion
everything not accounted for by the solid and NCGs, more than
likely, leading to a slight overestimation of the total liquid yield. The 4.1. Catalyst chemistry
total amount of condensable pyrolytic products decreases for each
catalyst when compared to non-catalytic reactions, a drawback In general, the acid catalyzed formation of aromatic hydrocar-
reported for catalytic pyrolysis of biomass [5,8]. NCG production bons begins with donation of a proton to incoming substrates such
is increased for each catalyst run compared to non-catalytic but as hydrocarbons at the acidic sites of the catalyst. This protonation
there does not appear to be a straightforward trend among the H- leads to carbocation formation within the hydrocarbon which then
ZSM-5 catalyst series. The overall NCG’s production was slightly reacts further by way of ␤-hydrogen elimination to form olefins
higher for H-ZSM-5 (50) at 32.7 wt% than for H-ZSM-5 (23) 30.9 wt% and the sequence continues. The olefins are transformed into aro-
while total NCG yields from H-ZSM-5 (280) were 25.2 wt%. This matics through oligomerization, cyclization, and hydrogen transfer
trend indicates the relative amount of aromatics formation cannot reactions also performed on the catalyst acid sites [43–45].
be directly related to formation of gases. Solid formation is fairly Zeolite catalyzed fast pyrolysis of biomass carbohydrates includ-
consistent among the three H-ZSM-5 and all are slightly increased ing glucose, cellobiose, cellulose, and xylitol as feedstocks in a
compared to non-catalytic. This consistency indicates that for the pyroprobe was described by Huber, et al. [46] The general pathway
H-ZSM-5 catalysts, coking is not dependent upon relative acidi- they established was that cellulose undergoes initial dehydration
230 D.J. Mihalcik et al. / Journal of Analytical and Applied Pyrolysis 92 (2011) 224–232

reactions to form anhydro-sugars which then interact with the leading to deoxygenation of ketones proceeds through decar-
active sites of the catalyst to form dehydrated products via acid cat- bonylation as was previously suggested by Adjaye & Bakhshi
alyzed dehydration. The dehydrated products further undergo acid [50].
catalyzed oligomerization, decarboxylation, and decarbonylation,
or cracking reactions, to form C2 -C6 olefins which then combine 4.2. Effects of catalyst acidity and framework
to yield aromatics. The deoxygenation occurs by the production
of CO, CO2 , and H2 O. Consistent with this pathway, we observed The higher acidity of the catalyst should theoretically increase
decreases in the production of oxygenates in general when zeolite the propensity of the catalyst to promote cracking reactions, but
catalysts were applied in the pyrolysis of cellulose and hemicellu- in the case of H-Y (5.1) zeolite, leads to the relative instability of
loses. Lignin pyrolysis over the catalysts studied followed pathway the catalyst framework. This instability led to catalyst framework
similar to that described by Mullen and Boateng where similar collapse and consequently, full compositional analysis of pyrolytic
catalysts specifically H-ZSM-5 and the mixed metal oxide catalyst products when using H-Y (5.1) was not performed. The low Si/Al
CoO/MoO3 were applied [42]. This resulted in the production of aro- ratio of this catalyst implies a relatively high acidity, where the
matic products which further undergo cracking reactions to olefins Al-O bond is susceptible to hydrolysis reactions which lead to the
leading to production of deoxygenated aromatics. breaking of Si-O-Al bonds and removal of the aluminum from the
Furthermore, upon pyrolysis of biomass components over Mor- tetrahedral lattice position further leading to lattice defects and
denite and Ferrierite catalyst, significant coke depositions were ultimately framework destabilization and eventual collapse [51].
not observed. For lignocellulosic biomass samples, there was an This mechanism could potentially lead to unreliable solid recovery
increase in solid residue formation in each biomass/catalyst com- values and as a result, wt% of condensables could not be calcu-
bination with the exception of the combination of corn cob and lated. This susceptibility to destabilization may also help to explain
H-Ferrierite. Because corn cob contains very low levels of lignin the lack of aromatics production despite the relatively favorable
this data suggest that lignin concentration is a contributing fac- acidity.
tor in the increased solid formations. Therefore, interactions of Regardless of relative acidity, the pore structure of the indi-
lignin in cellulosic biomass with H-Ferrierite are different than vidual catalyst plays a very important role. For instance, the H-Y
those interactions in pure lignin leading to increased coke for- (5.1) catalyst may be described as a cubic structure and is com-
mation in the former. Another potential factor is the presence of posed of straight channels measuring 0.74 nm in diameter defined
non-methoxylated lignins (H-lignin) which are more highly con- by a 12-member oxygen ring (0.74 nm); pores in perpendicular x, y,
centrated in herbaceous species (i.e. corn stover and switchgrass) and z planes. The potential problem with this type of pore system
than hardwoods (i.e. oak). may be two-fold. Firstly, the size of the channels is relatively large
The entire set of catalysts exhibit similar reactivity trends compared to other zeolites tested. This large size may potentially
towards deoxygenation of pyrolytic vapors, regardless of the feed- result in far less contact with incoming pyrolytic substrates and
stock and despite their different frameworks. This trend suggests therefore offers significantly less potential for cracking reactions
that they are acting on initial pyrolytic degradation products rather to occur. Secondly, the large and straight channels also contribute
than the carbohydrates or lignin components. The pyrolytic degra- to the ineffective contact ability but also shape selectivity becomes
dation products are then deoxygenated over the zeolite catalysts a non-issue. Essentially, the pyrolytic products may be channeling
to form the aromatic hydrocarbons. Adjaye and Bakhshi proposed through the catalyst relatively freely.
that the light oxygenated organics are converted to olefins which The pore size and framework of the catalyst tend to affect
then aromatize (Scheme 1) [47]. the product composition. For instance, the shape selectivity often
There are three types of pathway mechanisms for the deoxy- displayed by H-ZSM-5 catalysts [20] was observed in these experi-
genation of the oxygenated organics; dehydration to form water, ments. The H-ZSM-5 (23) seems to offer shape selectivity in terms of
decarbonylation to form CO, and decarboxylation to form CO2 . favorably cracking pyrolytic vapors to para rather than ortho-xylene
For products of cellulose pyrolysis, the increase in CO2 observed for all feedstocks tested. o-Xylene has a molecular dimension of
over the H-ZSM-5 suggests that both dehydration and decarboxyla- approximately 0.55 nm which is slightly larger than the average
tion mechanisms are active. Further, both Mordenite and Ferrierite width of H-ZSM-5 pores (0.52–0.55 nm). Whereas p-xylene has a
greatly reduced the production of levoglucosan from cellulose, with width closer to that of benzene at 0.51 nm, so it may be more
a qualitative, and co-concurrent, increase in production of substi- effectively produced and transported within the pore system of the
tuted furans when compared to non-catalytic reactions. This trend H-ZSM-5 catalyst.
indicates the catalytic pyrolysis of cellulose most likely proceeds Trends indicate that when considering the reduction of oxy-
through fragmentation, rather than depolymerization, as is the case genates, the H-Beta series and especially H-Beta (25) seem
with non-catalytic pyrolysis, but does not indicate that the con- to be more feedstock specific when compared to the H-ZSM-
densable fraction of the pyrolysis products was deoxygenated. The 5 series. For instance, even with the highest acidity, H-Beta
presence of these catalysts promotes pathways similar to those (25) was actually the least effective for deoxygenation of cel-
caused by biomass ash, i.e., and the presence of alkali metals lulose pyrolytic vapors and behaved similarly to the less acidic
reported to shift degradation mechanisms towards fragmentation H-Beta (38) for hemicellulose, pure lignin, and ETEK lignin.
[48]. Levoglucosan is deoxygenated in two steps, one dehydra- Interestingly, for oak, corn cob, and corn stover, a dramatic
tion to form furans and similar compounds and then conversion of difference is observed between the most acidic, and more
the furans to aromatic hydrocarbons via further dehydration and effective H-Beta (25), and its less acidic counterparts. This indi-
decarboxylation (Scheme 2). A possible mechanism for the con- cates that when using the H-Beta catalyst, the differing Si/Al
version of furans is acid catalyzed ring opening to form aldehydes ratios play a larger role than for lignocellulosic biomass feed-
(Scheme 3) which can be converted to aromatics by mechanisms stocks than for the component feedstocks. This discrepancy
similar to that proposed by Resasco for conversion of propanal between feedstocks did not pertain to the H-ZSM-5 catalyst
[49]. series.
For hemicellulose the major net reaction observed between The feedstock specificity observed for the H-Beta series for the
non-catalytic pyrolysis and catalytic pyrolysis over HZSM-5 was production of oxygenates was not reciprocated in the production of
conversion of acetol (a ketone) to aromatics, with an increased aromatics. While production of aromatics from H-ZSM-5 plateaus,
production of carbon monoxide. This suggests that the conversion production of aromatics using the H-Beta catalyst series is clearly
D.J. Mihalcik et al. / Journal of Analytical and Applied Pyrolysis 92 (2011) 224–232 231

Scheme 1. General pathways towards aromatic hydrocarbons from oxygenated pyrolytic substrates.

Scheme 2. Two-step deoxygenation of levoglucosan to form furans and aromatic hydrocarbons via further dehydration and decarboxylation.

Scheme 3. Proposed acid catalyzed conversion of aldehydes to aromatics.

dependent upon the overall acidity of the catalyst. For each compo- Acknowledgement
nent and lignocellulosic biomass feedstock, H-Beta (25) produces
more aromatics than other less acidic H-Beta zeolite catalyst. This Author Mihalcik thanks the Agricultural Research Service for
linear effect more closely describes the effect of decreasing acid partial funding through the Administrator Funded Research Asso-
strength on the production of aromatics and also exemplifies the ciate Program. We would like to thank statistician John G. Phillips
differences observed between types of catalyst. of the North Atlantic Area (NAA) for analyzing data by analysis of
H-Beta was affected by increased amounts of coke regardless variance to determine the effects and interactions of catalyst and
of its relative acidity. Coke is the carbonaceous solid result- feedstock.
ing from the decomposition or condensation of hydrocarbons
and is typically composed of polymerized heavy hydrocarbons. Appendix A. Supplementary data
Because of the increased amount of hydrocarbons being pro-
duced within the pores of the zeolites during catalytic pyrolysis, Supplementary data associated with this article can be found, in
there is an increased likelihood of hydrocarbon decomposi- the online version, at doi:10.1016/j.jaap.2011.06.001.
tion or condensation leading to coke formation within the
pores. The formation of coke within the zeolite pore system References
leads to fouling due to blockage and eventual catalyst deac-
tivation. Zeolite catalyst deactivation is of major concern and [1] G.W. Huber, S. Iborra, A. Corma, Synthesis of transportation fuels from
biomass: chemistry, catalysts, and engineering, Chemical Reviews 106 (2006)
presents reactor engineering challenges necessary for catalyst
4044–4098.
regeneration. [2] D. Mohan, C.U. Pittman Jr., P.H. Steele, Pyrolysis of wood/biomass for bio-oil: a
critical review, Energy and Fuels 20 (2006) 848–889.
[3] M.M. Wright, R.C. Brown, A.A. Boateng, Distributed processing of biomass to
5. Conclusion bio-oil for subsequent production of Fischer-Tropsch liquids, Biofuels Bioprod-
ucts Biorefining 2 (2008) 229–238.
[4] S. Schwietzke, M. Ladisch, L. Russo, K. Kwant, T. Makinen, B. Kavalov, K. Mani-
Eight lignocellulosic biomass and component feedstocks were atia, R. Zwart, G. Shananan, K. Sipila, P. Grabowwski, B. Telenius, M. White, A.
screened non-catalytically and catalytically using nine com- Brown, Report of the IEA Bioenergy Task 41, Project 2, 2008.
mercially available acidic zeolite catalysts through Py-GC/MS [5] P.T. Williams, N. Nugranad, Comparison of products from the pyrolysis and
catalytic pyrolysis of rice husks, Energy 25 (2000) 493–513.
experiments. Quantitative analysis of eight oxygenates and seven [6] J. Adam, M. Blazso, E. Meszaros, M. Stocker, M.H. Nilsen, A. Bouzga, J.E. Hustad,
aromatic hydrocarbons common to non-catalytic and catalytic M. Gronli, G. Oye, Pyrolysis of biomass in the presence of Al-MCM-41 type
pyrolysis was presented. Relative compositions of pyrolytic prod- catalysts, Fuel 84 (2005) 1494–1502.
[7] M.A. Jackson, D.L. Compton, A.A. Boateng, Screening heterogeneous catalysts
ucts: NCG, solid residues, and condensable vapors, were presented for the pyrolysis of lignin, Journal of Analytical and Applied Pyrolysis 85 (2009)
to show how reaction products are affected by catalyst choice. 226–230.
While all catalysts decreased the oxygen containing species in the [8] H. Zhang, R. Xaio, H. Huang, G. Xiao, Comparison of non-catalytic and catalytic
fast pyrolysis of corncob in a fluidized bed reactor, Bioresource Technology 100
pyrolytic vapors, the H-ZSM-5 (23) catalyst was the most effective (2009) 1428–1434.
catalyst at producing aromatic hydrocarbons from the oxygen-rich [9] A. Pattiya, J.O. Titiloye, A.V. Bridgwater, Fast pyrolysis of cassava rhizome in
vapors. The framework and Si/Al ratio of the catalysts plays a major the presence of catalysts, Journal of Analytical and Applied Pyrolysis 81 (2008)
72–79.
role in their ability to effectively deoxygenate the vapors and pro-
[10] E.F. Iliopoulou, E.V. Antonakou, S.A. Karakoulia, I.A. Vasalos, A.A. Lappas, K.S.
duce aromatic hydrocarbons. Triantafyllidis, Catalytic conversion of biomass pyrolysis products by meso-
232 D.J. Mihalcik et al. / Journal of Analytical and Applied Pyrolysis 92 (2011) 224–232

porous materials: effect of steam stability and acidity of Al-MCM-41 catalysts, [31] H. Zhang, R. Xiao, D. Wang, Z. Zhong, M. Song, Q. Pan, G. He, Catalytic fast
Chemical Engineering Journal 134 (2007) 51–57. pyrolysis of biomass in a fluidized bed with fresh and spent fluidized catalytic
[11] J. Adam, E. Antonakou, A. Lappas, M. Stocker, M.H. Nilsen, A. Bouzga, J.E. Hustad, cracking (FCC) catalysts, Energy Fuels 23 (2009) 6199–6206.
G. Oye, In situ catalytic upgrading of biomass derived fast pyrolysis vapors in a [32] H. Li, P. Yu, B. Shen, Biofuel potential production from cottonseed oil: a com-
fixed bed reactor using mesoporous materials, Microporous and Mesoporous parison of non-catalytic and catalytic pyrolysis on fixed-fluidizing bed reactor,
Materials 96 (2006) 93–101. Fuel Processing Technology 90 (2009) 1087–1092.
[12] A. Pattiya, J.O. Titiloye, A.V. Bridgwater, Evaluation of catalytic pyrolysis of [33] T.R. Carlson, T.P. Vispute, G.W. Huber, Green gasoline by catalytic fast
cassava rhizome by principle component analysis, Fuel 89 (2010) 244–253. pyrolysis of solid biomass derived compounds, ChemSusChem 1 (2008)
[13] E. Antonakou, A. Lappas, M.H. Nilsen, A. Bouzga, M. Stocker, Evaluation of var- 397–400.
ious types of Al-MCM-41 materials as catalysts in biomass pyrolysis for the [34] A. Atutxa, R. Aguado, A.G. Gayubo, M. Olazar, J. Bilbao, Kinetic description of
production of bio-fuels and chemicals, Fuel 85 (2006) 2202–2212. the catalytic pyrolysis of biomass in a conical spouted bed reactor, Energy Fuel
[14] D. Fabbri, C. Torri, V. Baravelli, Effect of zeolites and nanopowder metal oxides 19 (2005) 765–774.
on the distribution of chiral anhydrosugars evolved from pyrolysis of cellulose: [35] A. Aho, N. Kumar, K. Eranen, T. Salmi, M. Hupa, D.Y. Murzin, Fast pyrolysis
an analytical study, Journal of Analytical and Applied Pyrolysis 80 (2007) 24–29. of biomass for bio-oil with ionic liquid and microwave radiation, I Chemi-
[15] M.I. Nokkosmaki, E.T. Kuoppala, E.A. Leppamaki, A.O.I. Krause, Catalytic con- cal Engineering Part B Process Safety and Environmental Protection 85 (2007)
version of biomass pyrolysis vapors with zinc oxide, Journal of Analytical and 473–480.
Applied Pyrolysis 55 (2000) 119–131. [36] P.A. Horne, P.T. Williams, Upgrading of biomass-derived pyrolytic vapors over
[16] Q. Lu, W.-M. Xiong, W.-Z. Li, Q.-X. Guo, X.-F. Zhu, Catalytic pyrolysis of cellulose zeolite ZSM-5 catalyst: effect of catalyst dilution on product yields, Fuel 75
with sulfated metal oxides: a promising method for obtaining high yield of light (1996) 1043–1050.
furan compounds, Bioresource Technology 100 (2009) 4871–4876. [37] A.A. Lappas, M.C. Samolada, D.K. Latridis, S.S. Voutetakis, I.A. Vasalos, Biomass
[17] S.J. Juutilainen, P.A. Simell, A. Outi, I. Krause, Zirconia: selective oxidation cat- pyrolysis in a circulating fluid bed reactor for the production of fuels and chem-
alyst for removal of tar and ammonia from biomass gasification gas, Applied icals, Fuel 81 (2002) 2087–2095.
Catalysis B: Environmental 62 (2006) 86–92. [38] M. Kaylen, D.L. Van Dyne, Y.S. Choi, M. Blasé, Economic feasibility of produc-
[18] J. Li, B. Xiao, L. Du, Y. Rong, T.D. Liang, Formation of NiO nanowires on the surface ing ethanol from lignocellulosic feedstocks, Bioresource Technology 72 (2000)
of nickel strips, Journal of Fuel Chemistry and Technology 36 (2008) 42–47. 19–32.
[19] J. Li, R. Yan, B. Xiao, D.T. Liang, D.H. Lee, Preparation of nano-NiO particles and [39] A. Dermibas, Mechanisms of liquefaction and pyrolysis reactions of biomass,
evaluation of their catalytic activity in pyrolyzing biomass components, Energy Energy Conversion and Management 41 (2000) 633–646.
and Fuels 22 (2008) 16–23. [40] R.J. Evans, T.A. Milne, Molecular characterization of the pyrolysis of biomass,
[20] P.B. Weisz, W.O. Haag, P.G. Rodewald, Catalytic production of high-grade fuel Energy and Fuels 1 (1987) 123–137.
(gasoline) from biomass compounds by shape-selective catalysis, Science 206 [41] F. Schafizadeh, Introduction to pyrolysis of biomass, Journal of Analytical and
(1979) 57–58. Applied Pyrolysis 3 (1982) 283–305.
[21] M.I. Haniff, L.H. Dao, Deoxygenation of carbohydrates and their isopropylidene [42] C.A. Mullen, A.A. Boateng, Characterization of water insoluble solids isolated
derivatives over ZSM-5 zeolite catalysts, Applied Catalysis 39 (1988) 33–47. from various biomass fast pyrolysis oils, Fuel Processing Technology 91 (2010)
[22] M. Olazar, R. Aguado, J. Bilbao, Pyrolysis of sawdust in a conical spouted- 1446–1458.
bed reactor with a HZSM-5 catalyst, Reactors Kinetics and Catalysis 5 (2000) [43] D.B. Lukyanov, N.S. Gnep, M. Guisnet, Kinetic modeling of ethane and propene
1025–1033. aromatization over HZSM-5 and GaHZSM-5, Industrial and Engineering Chem-
[23] P.A. Horne, N. Nugranad, P.T. Williams, Catalytic coprocessing of biomass- istry Research 33 (1994) 223–234.
derived pyrolysis vapors and methanol, Journal of Analytical and Applied [44] Y. Song, X. Zhu, S. Xie, Q. Wang, L. Xu, The effect of acidity on olefin aroma-
Pyrolysis 34 (1995) 87–108. tization over potassium modified ZSM-5 catalysts, Catalysis Letters 97 (2004)
[24] R. French, S. Czernik, Catalytic pyrolysis of biomass for biofuels production, Fuel 31–36.
Processing Technology 91 (2010) 25–32. [45] V.R. Choudhary, D. Panjala, S. Banerjee, Aromatization of propene and n-
[25] A.A. Lappas, M.C. Samolada, D.K. Iatridis, S.S. Voutetakis, I.A. Vasaloa, Biomass butene over H-galloaluminosilicate (ZSM-5 type) zeolite, Applied Catalysis A
pyrolysis in circulating fluid bed reactor for the production of fuels and chem- 231 (2002) 243–251.
icals, Fuel (2002) 2087–2095. [46] T.R. Carlson, G.A. Tompsett, W.C. Conner, G.W. Huber, Aromatic production
[26] H.J. Park, J.-I. Dong, J.-K. Jeon, K.-S. Yoo, J.-H. Yim, J.-M. Sohn, Y.-K. Park, Con- from catalytic fast pyrolysis of biomass-derived feedstocks, Topics in Catalysis
version of the pyrolytic vapor of radiate pine over zeolites, Journal of Industrial 52 (2009) 241–252.
Chemistry 13 (2007) 182–189. [47] J.D. Adjaye, N.N. Bakhshi, Production of hydrocarbons by catalytic upgrading of
[27] A. Aho, N. Kumar, K. Eranen, T. Salmi, M. Hupa, D.Y. Murzin, Catalytic pyrolysis a fast pyrolysis bio-oil.Part II: comparative catalyst performance and reaction
of biomass in a fluidized bed reactor: influence of the acidity of H-Beta zeolite, pathways, Fuel Processing Technology 45 (1995) 185–202.
Process Safety and Environmental Protection 85 (2007) 473–480. [48] R. Fahmi, A.V. Bridgwater, L.I. Darvell, J.M. Jones, N. Yates, S. Thain, I.S. Donnison,
[28] A. Aho, N. Kumar, K. Eranen, T. Salmi, M. Hupa, D.Y. Murzin, Catalytic pyrolysis The effect of alkali metals on combustion and pyrolysis of Lolium and Festuca
of woody biomass in a fluidized bed reactor: influence of the zeolite structure, grasses, switchgrass, and willow, Fuel 86 (2007) 1560–1569.
Fuel 87 (2008) 2493–2501. [49] T.Q Hoang, X. Zhu, T. Sooknoi, D.E. Resasco, R.G. Mallinson, A comparison of
[29] R.T. Carlson, G.A. Tompsett, W.C. Conner, G.W. Huber, Aromatic production the reactivities of propanal and propylene on HZSM-5, Journal of Catalysis 271
from catalytic fast pyrolysis of biomass-derived feedstocks, Topics in Catalysis (2010) 201–208.
52 (2009) 241–252. [50] J.D. Adjaye, N.N. Bakhshi, Catalytic conversion of a biomass-derived oil to fuels
[30] E.M. Sulman, V.V. Alferov, Y.Y. Kosivtsov, A.I. Sidorov, O.S. Misnikov, A.E. and chemicals I: model compound studies and reaction pathways, Biomass and
Afanasiev, N. Kumar, D. Kubicka, J. Agullo, T. Salmi, D.Y. Murzin, The devel- Bioenergy 8 (1995) 131–149.
opment of the method of low-temperature peat pyrolysis on the basis of [51] L. Moscou, R. Mone, Structure and catalytic properties of thermally and
aluminosilicate catalytic system, Chemical Engineering Journal 134 (2007) hydrothermally treated zeolites: acid strength distribution of REX and REY,
162–167. Journal of Catalysis 30 (1973) 417–422.
Supplementary Information:

Table S1: Quantitative and statistical analysis of the wt% of pyrolytic products (1-15) observed
following non-catalytic pyrolysis of component and lignocellulosic feedstocks, and relative
values of NCG’s, solid residue, and total calculated condensable vapors.a
Cellulose Hemi- Lignin ETEK Oak Corn Cob Corn Stover Switchgrass
Cellulose Lignin
Acetic acid 1.89C 3.19C 2.21C 2.11C 7.25AB 8.19A 7.11AB 6.52B
Furfural 1.26AB 1.33AB 0.81C 0.96BC 1.42A 1.19ABC 1.19ABC 1.41A
Furfural Alcohol 0.17E 0.18DE 0.19DE 0.20D 0.28BC 0.31A 0.25C 0.30AB
Acetol 0.85EF 7.06A 0.10F 2.10D 1.86DE 5.83B 3.51C 5.26B
Levoglucosan 61.50A 0.0073E 0.20DE 12.61B 5.60C 0.81DE 2.21D 0.72DE
Guaiacol 0.015E 0.075DE 0.49A 0.51A 0.19BC 0.28B 0.17CD 0.44A
Syringol 0.018D 0.025D 1.01A 0.016D 0.33B 0.15C 0.16C 0.27B
Phenol 0.095C 0.12BC 0.20AB 0.14BC 0.098C 0.25A 0.16BC 0.26A
Benzene 0.16A 0.17A 0.17A 0.17A 0.14A 0.17A 0.17A 0.19A
Toluene 0.11A 0.18A 0.16A 0.14A 0.14A 0.19A 0.16A 0.21A
Ethyl benzene 0.012A 0.13A 0.13A 0.016A 0.010A 0.13A 0.019A 0.16A
p-xylene 0.015A 0.16A 0.14A 0.13A 0.014A 0.14A 0.13A 0.14A
o-xylene 0.0053B 0.14AB 0.14AB 0.018B 0.0068B 0.29A 0.019B 0.15AB
Naphthalene 0.012A 0.018A 0.011A 0.021A 0.014A 0.014A 0.013A 0.065A
Methyl Naphthalene 0.0085A 0.018A 0.010A 0.011A 0.0061A 0.024A 0.0063A 0.032A
NCG’s 11.7 21.7 8.8 7.3 14.7 21.2 13.8 13.8
Solid Residue 1.3 21.6 20.3 20.4 16.2 14.8 15.5 16.1
Condensables 87.0 56.7 70.9 72.3 69.1 64.0 70.7 70.1
a
Any two means in the same row with no letter in common are significantly different (p < 0.05) by the Bonferroni least significant difference
(LSD) technique. The letters A-F refer to the highest estimates to the least.
Table S2: Quantitative and statistical analysis of the wt% of pyrolytic products (1-15) observed
following catalytic pyrolysis of component and lignocellulosic feedstocks over H-Ferrierite (20)
and relative values of NCG’s, solid residue, and total calculated condensable vapors.a
Cellulose Hemi- Lignin ETEK Oak Corn Cob Corn Stover Switchgrass
Cellulose Lignin
Acetic acid 1.88A 2.18A 0.0079B 1.88A 2.15A 3.23A 2.68A 2.67A
Furfural 1.97A 0.80C 0.021D 0.88C 1.46B 1.19BC 1.20BC 0.015D
Furfural Alcohol 0.013A 0.016A 0.017A 0.018A 0.012A 0.012A 0.022A 0.011A
Acetol 0.23B 1.03AB 0.017B 0.42AB 0.73AB 1.48A 1.11AB 0.24B
Levoglucosan 1.82A 0.0060A 0.014A 0.017A 0.44A 0.022A 0.014A 0.010A
Guaiacol 0.012D 0.016D 0.51B 0.65A 0.28C 0.26C 0.24C 0.096D
Syringol 0.012E 0.013E 1.03A 0.015E 0.50B 0.094D 0.20C 0.018E
Phenol 0.13BC 0.13BC 0.20AB 0.20AB 0.13BC 0.26A 0.26A 0.11C
Benzene 0.22A 0.28A 0.30A 0.23A 0.23A 0.37A 0.25A 0.56A
Toluene 0.22A 0.60A 0.41A 0.33A 0.33A 0.57A 0.35A 0.68A
Ethyl benzene 0.15BC 0.34A 0.31AB 0.14BC 0.12C 0.19ABC 0.17ABC 0.34A
p-xylene 0.15A 0.38A 0.35A 0.19A 0.19A 0.30A 0.19A 0.38A
o-xylene 0.016B 0.13AB 0.13AB 0.13AB 0.0047B 0.14AB 0.079AB 0.22A
Naphthalene 0.10A 0.10A 0.097A 0.11A 0.10A 0.11A 0.11A 0.23A
Methyl Naphthalene 0.017A 0.11A 0.045A 0.049A 0.043A 0.055A 0.047A 0.11A
NCG’s 11.9 31.5 13.9 15.9 23.4 22.6 18.0 23.1
Solid Residue 1.7 18.3 20.0 16.7 20.3 13.8 28.2 31.2
Condensables 86.4 50.2 66.1 67.4 56.3 63.6 53.8 45.7
a
Any two means in the same row with no letter in common are significantly different (p < 0.05) by the Bonferroni least significant difference
(LSD) technique. The letters A-F refer to the highest estimates to the least.
Table S3: Quantitative and statistical analysis of the wt% of pyrolytic products (1-15) observed
following catalytic pyrolysis of component and lignocellulosic feedstocks over H-Mordenite (20)
and relative values of NCG’s, solid residue, and total calculated condensable vapors.a
Cellulose Hemi- Lignin ETEK Oak Corn Cob Corn Switchgrass
Cellulose Lignin Stover
Acetic acid 2.41BC 2.13BC 0.018D 1.81C 2.61BC 3.51AB 2.15BC 4.12A
Furfural 1.16A 0.59B 0.0056C 0.72B 0.79AB 0.011C 0.0050C 0.83AB
Furfural Alcohol 0.016A 0.023A 0.020A 0.022A 0.0069A 0.015A 0.0097A 0.019A
Acetol 0.19A 0.70A 0.0096A 0.55A 0.34A 0.92A 0.15A 0.53A
Levoglucosan 0.48A 0.017A 0.015A 0.021A 0.010A 0.014A 0.010A 0.40A
Guaiacol 0.020D 0.014D 0.22AB 0.26A 0.12BCD 0.11BCD 0.084CD 0.17ABC
Syringol 0.011C 0.016C 0.22A 0.0096C 0.095B 0.035C 0.019C 0.056BC
Phenol 0.16BC 0.097C 0.21B 0.31A 0.18B 0.17BC 0.16BC 0.30A
Benzene 0.86A 0.74AB 0.46B 0.70AB 0.72AB 0.84A 0.81AB 0.73AB
Toluene 0.24B 0.91A 0.40AB 0.51AB 0.71AB 0.81A 0.83A 0.76AB
Ethyl benzene 0.35AB 0.41A 0.20BC 0.16C 0.16C 0.17C 0.17BC 0.27ABC
p-xylene 0.24A 0.47A 0.21A 0.38A 0.37A 0.38A 0.41A 0.20A
o-xylene 0.14A 0.23A 0.16A 0.20A 0.21A 0.21A 0.22A 0.16A
Naphthalene 0.018B 0.22A 0.14AB 0.17AB 0.18AB 0.23A 0.20A 0.017B
Methyl Naphthalene 0.0043A 0.11A 0.066A 0.096A 0.088A 0.11A 0.099A 0.081A
NCG’s 14.1 18.7 14.6 13.9 16.5 22.3 19.4 21.7
Solid Residue 3.9 19.9 21.8 18.3 19.1 19.5 23.0 24.2
Condensables 82.0 61.4 63.6 67.8 64.4 58.2 57.6 54.1
a
Any two means in the same row with no letter in common are significantly different (p < 0.05) by the Bonferroni least significant difference
(LSD) technique. The letters A-F refer to the highest estimates to the least.
Table S4: Quantitative and statistical analysis of the wt% of pyrolytic products (1-15) observed
following non-catalytic and catalytic pyrolysis of component and lignocellulosic feedstocks.
Data arranged to allow for direct comparison of effects of catalyst/feedstock combinations on
individual condensable pyrolysis products (1-15).a
Cellulose Hemi- Lignin ETEK Oak Corn Cob Corn Switchgrass
Cellulose Lignin Stover

Acetic Acid
None 1.89A 3.19A 2.21A 2.11A 7.25A 8.19A 7.11A 6.52A
H-Ferrierite (20) 1.88A 2.18A 0.0079B 1.88A 2.15B 3.23B 2.68B 2.67C
H-Mordenite (20) 2.41A 2.13A 0.018B 1.81A 2.61B 3.51B 2.15B 4.12B
H-Y (5.1) 0.0093B 1.89A 0.020B 2.13A 0.0073C 3.25B 0.059C 4.33B
H-ZSM-5 (23) 2.25A 2.02A 0.0091B 1.72A 2.31B 3.07BC 1.72B 2.17C
H-Beta (25) 2.01A 1.92A 0.017B 1.75A 0.0082C 1.76C 1.90B 2.23C

Furfural
None 1.23B 1.33A 0.81A 0.96A 1.42A 1.19A 1.19A 1.41A
H-Ferrierite (20) 1.97A 0.80B 0.021B 0.88A 1.46A 1.19A 1.20A 0.015C
H-Mordenite (20) 1.16BC 0.59B 0.0056B 0.72A 0.79B 0.011B 0.0050B 0.83B
H-Y (5.1) 1.44B 0.017C 0.021B 1.04A 1.45A 0.99A 0.98A 1.16AB
H-ZSM-5 (23) 0.76CD 0.0079C 0.022B 0.011B 0.016C 0.013B 0.014B 0.020C
H-Beta (25) 0.68D 0.018C 0.018B 0.016B 0.025C 0.016B 0.013B 0.23C

Furfural Alcohol
None 0.17A 0.18A 0.19A 0.20A 0.28A 0.31A 0.25A 0.30A
H-Ferrierite (20) 0.013B 0.016B 0.017B 0.014B 0.012B 0.012B 0.022B 0.011B
H-Mordenite (20) 0.016B 0.023B 0.020B 0.022B 0.023B 0.015B 0.0097B 0.019B
H-Y (5.1) 0.017B 0.011B 0.0065B 0.017B 0.025B 0.018B 0.011B 0.020B
H-ZSM-5 (23) 0.019B 0.013B 0.018B 0.010B 0.023B 0.026B 0.017B 0.010B
H-Beta (25) 0.020B 0.0080B 0.021B 0.018B 0.017B 0.0021B 0.012B 0.016B

Acetol
None 0.85A 7.06A 0.10A 2.10A 1.86A 5.83A 3.51A 5.26A
H-Ferrierite (20) 0.23A 1.03B 0.017A 0.42B 0.73AB 1.48B 1.11BC 0.24C
H-Mordenite (20) 0.19A 0.70B 0.0096A 0.55B 0.34B 0.92BCD 0.15C 0.53C
H-Y (5.1) 0.35A 0.39B 0.013A 1.20AB 1.16AB 1.33BC 1.48B 2.22B
H-ZSM-5 (23) 0.12A 0.12B 0.010A 0.12B 0.93AB 0.31CD 0.013C 0.020C
H-Beta (25) 0.35A 0.58B 0.0046A 0.14B 0.0088B 0.12D 0.016C 0.11C

Levoglucosan
None 61.50A 0.0073A 0.20A 12.61A 5.60A 0.81A 2.21A 0.72A
H-Ferrierite (20) 1.82B 0.0060A 0.014A 0.017B 0.44B 0.022A 0.014B 0.010A
H-Mordenite (20) 0.48B 0.017A 0.015A 0.021B 0.010B 0.014A 0.010B 0.40A
H-Y (5.1) 1.85B 0.015A 0.011A 0.014B 1.10B 0.0068A 1.13AB 0.12A
H-ZSM-5 (23) 0.31B 0.022A 0.019A 0.019B 0.48B 0.0094A 0.025B 0.019A
H-Beta (25) 0.013B 0.012A 0.016A 0.018B 0.14B 0.0093A 0.018B 0.021A

Guaiacol
None 0.015A 0.075A 0.49A 0.51B 0.19AB 0.28A 0.17AB 0.44A
H-Ferrierite (20) 0.012A 0.016A 0.51A 0.65A 0.28A 0.26AB 0.24A 0.096C
H-Mordenite (20) 0.020A 0.014A 0.22BC 0.26D 0.12BC 0.11C 0.084B 0.17C
H-Y (5.1) 0.016A 0.020A 0.24B 0.39C 0.20AB 0.17BC 0.13AB 0.30B
H-ZSM-5 (23) 0.019A 0.010A 0.12C 0.15E 0.099BC 0.10C 0.084B 0.093C
H-Beta (25) 0.018A 0.021A 0.13C 0.11E 0.019C 0.080C 0.077B 0.073C

Syringol
None 0.018A 0.025A 1.01A 0.016A 0.33B 0.15A 0.16AB 0.27A
H-Ferrierite (20) 0.012A 0.013A 1.03A 0.015A 0.50A 0.094AB 0.20A 0.018C
H-Mordenite (20) 0.011A 0.016A 0.22C 0.0096A 0.095C 0.019C 0.019C 0.056C
H-Y (5.1) 0.010A 0.0084A 0.42B 0.013A 0.32B 0.035BC 0.12B 0.15B
H-ZSM-5 (23) 0.013A 0.023A 0.22C 0.018A 0.12C 0.048BC 0.037C 0.016C
H-Beta (25) 0.019A 0.0087A 0.063D 0.023A 0.014D 0.010C 0.016C 0.016C

Phenol
None 0.095A 0.12A 0.20A 0.14BC 0.098AB 0.25AB 0.16B 0.26A
H-Ferrierite (20) 0.13A 0.13A 0.20A 0.20B 0.13AB 0.26A 0.26A 0.11B
H-Mordenite (20) 0.16A 0.097A 0.21A 0.31A 0.18A 0.17BC 0.16B 0.30A
H-Y (5.1) 0.10A 0.097A 0.18A 0.16BC 0.13AB 0.18ABC 0.18B 0.29A
H-ZSM-5 (23) 0.14A 0.10A 0.15A 0.22B 0.16AB 0.19ABC 0.17B 0.11B
H-Beta (25) 0.10A 0.083A 0.16A 0.12C 0.087B 0.11C 0.11B 0.10B

Benzene
None 0.16C 0.17D 0.17B 0.17D 0.14D 0.17D 0.17D 0.19D
H-Ferrierite (20) 0.22C 0.28D 0.30B 0.23D 0.23D 0.37D 0.25D 0.56CD
H-Mordenite (20) 0.86B 0.74BC 0.46B 0.70C 0.72C 0.84C 0.81C 0.73BC
H-Y (5.1) 0.18C 0.38CD 0.45B 0.19D 0.24D 0.20D 0.35D 0.40CD
H-ZSM-5 (23) 3.06A 1.48A 1.39A 2.96A 2.93A 3.12A 3.17A 1.86A
H-Beta (25) 1.12B 0.84B 1.33A 1.48B 1.44B 1.51B 1.43B 0.96B

Toluene
None 0.11C 0.18D 0.16C 0.14B 0.14C 0.19C 0.16C 0.21C
H-Ferrierite (20) 0.22C 0.60CD 0.41C 0.33B 0.33BC 0.57BC 0.35BC 0.68C
H-Mordenite (20) 0.24C 0.91BC 0.40C 0.51B 0.71B 0.81B 0.83B 0.76C
H-Y (5.1) 0.24C 0.69CD 0.65C 0.19B 0.19BC 0.50BC 0.47BC 0.43C
H-ZSM-5 (23) 3.79A 2.99A 2.68A 2.24A 2.28A 2.23A 2.17A 3.36A
H-Beta (25) 1.47B 1.27B 1.83B 1.95A 2.05A 2.17A 2.21A 1.44B

Ethyl Benzene
None 0.012D 0.13C 0.13B 0.016B 0.010C 0.13B 0.019C 0.16D
H-Ferrierite (20) 0.15D 0.34B 0.31A 0.14AB 0.12BC 0.19B 0.17BC 0.34C
H-Mordenite (20) 0.35C 0.41B 0.20AB 0.16AB 0.16BC 0.17B 0.17BC 0.27CD
H-Y (5.1) 0.14D 0.29BC 0.29AB 0.13AB 0.14BC 0.14B 0.14BC 0.26CD
H-ZSM-5 (23) 1.77A 0.26BC 0.21AB 0.30A 0.37A 0.41A 0.37A 2.10A
H-Beta (25) 0.62B 0.62A 0.17AB 0.20A 0.20AB 0.25AB 0.26AB 0.66B

p-xylene
None 0.015C 0.16C 0.14C 0.13C 0.014D 0.14C 0.13C 0.14CD
H-Ferrierite (20) 0.15C 0.38C 0.35C 0.19C 0.19CD 0.30C 0.19C 0.38C
H-Mordenite (20) 0.24C 0.47BC 0.21C 0.38C 0.37C 0.38C 0.41C 0.20CD
H-Y (5.1) 0.15C 0.42BC 0.33C 0.17C 0.15CD 0.23C 0.22C 0.016D
H-ZSM-5 (23) 2.15A 2.59A 2.46A 3.17A 4.12A 3.49A 3.65A 2.56A
H-Beta (25) 0.73B 0.73B 0.96B 0.97B 1.10B 1.09B 1.21B 0.78B

o-xylene
None 0.0053C 0.14C 0.14B 0.018D 0.0068D 0.29BC 0.019D 0.15C
H-Ferrierite (20) 0.016C 0.13C 0.13B 0.13CD 0.0047D 0.14C 0.079CD 0.22BC
H-Mordenite (20) 0.14C 0.23BC 0.16B 0.20C 0.21C 0.21C 0.22C 0.16C
H-Y (5.1) 0.12C 0.21BC 0.16B 0.14CD 0.14CD 0.15C 0.16CD 0.15C
H-ZSM-5 (23) 0.54A 0.59A 0.40A 0.63A 0.74A 0.77A 0.77A 0.55A
H-Beta (25) 0.32B 0.31B 0.39A 0.38B 0.43B 0.43B 0.47B 0.33B

Naphthalene
None 0.012C 0.018D 0.011C 0.021C 0.014C 0.014D 0.013D 0.065D
H-Ferrierite (20) 0.10C 0.10CD 0.097BC 0.11C 0.10C 0.11CD 0.11CD 0.23C
H-Mordenite (20) 0.018C 0.22BC 0.14BC 0.17C 0.18C 0.23C 0.20C 0.017D
H-Y (5.1) 0.10C 0.16CD 0.19B 0.10C 0.11C 0.15CD 0.15CD 0.070CD
H-ZSM-5 (23) 1.44A 0.50A 0.75A 1.33A 1.37A 1.33A 1.33A 0.89A
H-Beta (25) 0.40B 0.38AB 0.70A 0.53B 0.60B 0.63B 0.56B 0.48B

1-methyl Naphthalene
None 0.0085C 0.014C 0.010C 0.011C 0.0061C 0.024C 0.0063C 0.032B
H-Ferrierite (20) 0.017C 0.11BC 0.045C 0.049C 0.043C 0.055C 0.047C 0.11B
H-Mordenite (20) 0.0043C 0.11BC 0.066C 0.096C 0.088C 0.11C 0.099C 0.081B
H-Y (5.1) 0.016C 0.060C 0.080C 0.038C 0.038C 0.060C 0.058C 0.041B
H-ZSM-5 (23) 0.85A 0.60A 0.70A 1.19A 1.28A 1.16A 1.12A 0.81A
H-Beta (25) 0.18B 0.25B 0.43B 0.30B 0.35B 0.39B 0.37B 0.15B
a
Any two means within same column of respective compound with no letter in common are significantly different (p < 0.05) by the Bonferroni
least significant difference (LSD) technique. The letters A-E refer to the highest estimates to the least.
Table S5: Quantitative and statistical analysis of the wt% of pyrolytic products (1-15) observed
following catalytic pyrolysis of component and lignocellulosic feedstocks over H-Y (5.1).a
Cellulose Hemi- Lignin ETEK Oak Corn Cob Corn Switchgrass
Cellulose Lignin Stover
C B C B C AB C
Acetic acid 0.0093 1.89 0.020 2.13 0.0073 3.25 0.0059 4.33A
Furfural 1.44A 0.017C 0.021C 1.04AB 1.45A 0.99B 0.98B 1.16AB
Furfural Alcohol 0.017A 0.011A 0.0065A 0.17A 0.025A 0.018A 0.011A 0.020A
Acetol 0.35BC 0.39BC 0.013C 1.20AB 1.16ABC 1.33AB 1.48AB 2.22A
Levoglucosan 1.85A 0.015A 0.011A 0.014A 1.10A 0.017A 1.13A 0.12A
Guaiacol 0.016D 0.020D 0.24BC 0.39A 0.20BC 0.17C 0.13C 0.30AB
Syringol 0.010D 0.0084D 0.42A 0.015D 0.32B 0.060D 0.12C 0.15C
Phenol 0.10CD 0.097D 0.18B 0.16BCD 0.13BCD 0.18BC 0.18BC 0.29A
Benzene 0.18A 0.38A 0.45A 0.19A 0.24A 0.20A 0.35A 0.40A
Toluene 0.24A 0.69A 0.65A 0.19A 0.19A 0.50A 0.47A 0.43A
Ethyl benzene 0.14A 0.29A 0.29A 0.13A 0.14A 0.14A 0.14A 0.26A
p-xylene 0.15A 0.42A 0.33AB 0.17AB 0.15AB 0.23AB 0.22AB 0.016B
o-xylene 0.12A 0.21A 0.16A 0.14A 0.14A 0.15A 0.16A 0.15A
Naphthalene 0.10A 0.16A 0.19A 0.10A 0.11A 0.15A 0.15A 0.070A
Methyl Naphthalene 0.016A 0.060A 0.080A 0.038A 0.038A 0.060A 0.058A 0.041A
a
Any two means in the same row with no letter in common are significantly different (p < 0.05) by the Bonferroni least significant difference
(LSD) technique. The letters A-F refer to the highest estimates to the least.

You might also like