Sie sind auf Seite 1von 11

JOURNAL 0F FOREST PRODUCTS AND INDUSTRIES, 2015, 4(2), 33-43 ISSN: 23254513(print) ISSN 2325 - 453X (Online)

Research Article

33

Catalytic Upgrading of Pinewood Fast Pyrolysis Vapors using an


Integrated Auger Packed Bed Reactor System: Effects of Acid
Catalysts on Yields and Distribution of Pyrolysis Products
Vamshi Krishna Guda1,* and Hossein Toghiani2

1, 2

Swalm School of Chemical Engineering, Mississippi State University, Mississippi State, MS-39762
*Corresponding author: vg97@msstate.edu

(Received: March 14, 2015; Accepted: May 12, 2015)

Abstract: In-situ catalytic upgrading of pine wood fast


pyrolysis vapors was performed using an integrated reactor
set-up where fast pyrolysis of pine wood chips was carried out in
an auger reactor followed by catalytic upgrading in a packed
bed reactor mounted on the top of auger reactor. The pyrolysis
process was carried out at 450 C under nitrogen atmosphere.
The effects of catalyst composition (-Al2O3, Si/Al, HY, HZSM5)
were studied in upgrading pine wood fast pyrolysis vapors.
Compared to bio-oil from non-catalytic experiments,
deoxygenated products increased in all the bio-oils obtained
from all the tested acid catalysts. While Si/Al and -Al2O3
yielded higher liquid products, the zeolites (HZSM5, HY)
favored the formation of aromatic hydrocarbons and higher
gaseous yields. HZSM5 catalyzed bio-oil was the most
deoxygenated bio-oil having lowest acid number (46.4%), lowest
oxygen content (30%) and high amount of phenolics among all
the catalyzed bio-oils.
Index terms fast pyrolysis; auger reactor; bio-oil; catalytic
cracking; in-situ upgrading.

I. INTRODUCTION
Depleting fossil fuel resources, combined with ever
increasing dependence on petroleum fuels through out the
world, and the political, economic, and environmental
concerns regarding fossil fuels initiated the societys quest for
alternative energy resources. A huge interest has been
developed for lignocellulosic biomass as an alternative
energy resource because of its vast abundancy, sustainability
and environmentally benign nature. Fast pyrolysis of
lignocellulosic biomass produces high energy density liquid
fuels, called as bio-oils, that have potential to act as
transportation fuels [1-3]. But, crude bio-oils are highly
oxygenated (40 % oxygen content), viscous, corrosive,
relatively unstable, and chemically complex liquids that have

low heating value and poor miscibility with petroleum fuels


[4, 5]. Therefore, the bio-oils have to be substantially
upgraded (deooxygenated) to highly stable, non-corrosive,
and high calorific value liquid fuels prior to their use as
transportation fuels.
Various upgrading techniques have been employed to convert
bio-oils to a variety of end products [6, 7] including hydrogen
[8-10], sugars [11], alkanes [9, 12], olefins [13], gasoline
range aliphatic and aromatic hydrocarbons [14], etc. Catalytic
processes such as aquaeous phase reforming [10, 12],
hydrodeoxygenation [15-17], and catalytic cracking [18] are
widely followed upgrading processes. Aqueous phase
reforming to convert the sugars and acids of bio-oil aqueous
fraction to hydrogen and other oxygenates [8]. The catalytic
upgrading of bio-oils has been mainly concentrated on
catalytic vapor cracking rather than hydrotreaing because of
low operational costs (no use of H2 and ambient reaction
pressures). Moreover, a wide variety of inexpensive cracking
catalysts can be used for catalytic vapor cracking either for
simple stabilization of bio-oils or to obtain high quality fuel
products. Catalytic cracking experiments were either
performed by condensing liquids from fast pyrolysis
experiments and upgrading the condensed bio-oils in a
secondary unit (ex-situ or secondary upgrading) [19-21] or by
directly treating the pyrolytic vapors coming out of pyrolysis
reactor [18, 22, 23] (catalytic pyrolysis or pyrolysis vapor
upgrading). In-situ upgrading eliminates the additional
bio-oil condensation and re-evoporation steps otherwise
required for ex-situ upgrading and therefore, is relatively cost
effective. Moreover, ex-situ upgrading of bio-oil will result in
loss of liquid products to coke (when heated to reaction
temperature) due to poor thermal stability of bio-oils.

34
JOURNAL 0F FOREST PRODUCTS AND INDUSTRIES, 2015, 4(2), 33-43 ISSN: 23254513(print) ISSN 2325 - 453X (Online)

Amorphous Si-Al catalyst, crystalline Si-Al catalysts such as


zeolites, and metal substituted zeolites are the most widely
employed cracking catalysts. While Si-Al catalysts produced
low yields of aliphatic compounds [19, 24], zeolites resulted
in significant yields of aromatic hydrocarbons [25]. Recently,
mesoporous aluminosilicate catalysts belonging to MCM
family (MCM-41) have been tested in upgrading the
pyrolysis vapors [26-30]. High phenolic yields were obtained
but the formation of polyaromatics (attributed to mesoporous
nature) and low hydrothermal stability of catalysts were the
drawbacks associated [31, 32]. Mesoprous MFI catalysts
were prepared by combining the properties of zeolites
(bronsted acidity, hydrothermal stabilty) and MCM
(mesoporous nature) family catalyts to obtain mesoporous
catalysts with higher acidity and hydrothermal stability than
MCM-41 catalysts. MFI zeolites, compared to MCM-41 and
ZSM5 catalysts, markedly enhanced the deoxygenation of
bio-oil while selectively producing monoaromatics, including
benzene, toluene, and xylene (BTXs), but resulted in poor
organic fraction yields [33, 34]. Zeolites, due to their
bronsted acidity and shape selective catalysis, are reported to
be the most effective catalysts, however associated with
drawbacks such as low organic product yields and rapid
deactivation of the catalysts. Recent studies [25] proved that
rapid heating rates and high catalyst to feed ratios attenuate
the catalyst deactivation while improving organic product
yields.
Catalytic pyrolysis studies were predominantly perfomed
using fluidized bed reactor [35-38]. Fluidized bed reactor
provides high heat transfer rates but require high N 2 flow to
achieve the required fluidization of biomass, catalyst, and
heat carrier particles. A highly efficient cyclone is required to
separate/ filter char particles from fast pyrolysis vapors
before the pyrolysis vapors are condensed to bio-oil.
Moreover, continuous feed/ recycling of the heat carrier (if
used) is required to counter the loss of carrier particles that
exit the reactor along with pyrolysis vapors. Fast pyrolysis
bio-oils produced using auger reactor were reported to be
comparable to bio-oil produced using fluidized bed [39-41].
Moreover, auger reactor is portable, suitable for large
biomass particles and does not require high inert gas flow or
a cyclone system. Data on catalytic upgrading of pyrolysis
vapors using an integrated system involving auger reactor
[42, 43] is scant and requires exploring further .
Therefore, the objective of this paper is to study catalytic
vapor upgrading of pine wood pyrolysis vapors using an

integrated reactor system where fast pyrolysis of pine wood


was perfomed using an auger reactor while the catalytic
treatment of pine wood pyrolysis vapors was achieved using a
packed bed reactor (PBR). The present studies concentrated
on testing the effects of four different acid catalysts (Si/Al,
Alumina, Zeolite-Y, and HZSM5) on the yields and
selectivity of pyrolysis products. .

II. MATERIALS AND METHODS


Feedstock
Pine wood tree stems were debarked and then reduced to
19-32 mm sized paper chips. The chips were ground to
smaller sized particles in a hammer mill and finally sieved to
the required particle sizes (2-3 mm) for the pyrolysis
experiments. The pine wood chips were oven dried at 105 oC
overnight to about 8-10% moisture content and stored in
sealed plastic buckets prior to use in the experiments.
Catalysts
Amorphous SiO2/Al2O3 (SiO2/Al2O3=5.15; Si/Al), -alumina
(99% alumina), Zeolite-Y (ZY; SiO2/Al2O3 =30), and
HZSM5 (SiO2/Al2O3 = 23) were used for upgrading pine
wood fast pyrolysis vapors. The zeolite supports, HZSM5 and
ZY, were commercially purchased from Zeolyst
International. Si/Al and -alumina catalysts were supplied by
Saint-Gobain for research purposes. The catalysts were
calcined in an oven at 550 oC for 8 hours prior to their use in
experiments. Catalyst properties are shown in Table 1.
Reactor Configuration and Experimental set-up
An auger reactor pipe of 7.6 cm in diameter and 102 cm in
length was employed for the purpose of producing bio-oil
vapor. The distance between the screws (i.e., between the
augers) is about 3 inches. The auger speed was 11-12 RPM at
the desired auger pyrolysis temperature. The particle size of
the pine sawdust used in the pyrolysis was ~2-3 mm.
According to the literature in which the same auger reactor
was employed [40], it takes approximately 45 sec for the
initial feed to move to the char exit point. The auger reactor
employed for bio-oil production has four temperature zones,
which facilitate the decomposition of different wood
constituents including cellulose, hemicellulose, and lignin, in
the appropriate temperature range, thereby forming liquid,
gaseous, and solid (char) products. The feed first passes
through a 4 heating (preheating) zone kept at ~200-220 C,

35
JOURNAL 0F FOREST PRODUCTS AND INDUSTRIES, 2015, 4(2), 33-43 ISSN: 23254513(print) ISSN 2325 - 453X (Online)

then enters two pyrolysis zones, 10 and 8 in length,


respectively, with temperatures in the range 450 C, and
finally reaches an 8 zone kept at ~150-175 C. Initial feed is
converted to char by the time it reaches the 8 fourth zone.
Upon exit, the char is collected in a collection pot through a
3 long insulated zone. The gases produced from biomass
pyrolysis then passed through a 1 diameter, 24 PBR
mounted on the top of the auger reactor. This PBR was
enclosed in a three zone furnace. The first two zones were
kept at the desired catalytic upgrading reaction temperatures
(425oC) whereas the third zone is maintained at a temperature
lower (by 50-75oC) than the first two to start cooling of
pyrolysis vapors before reaching the condensers. The effluent
pyrolysis vapors from the PBR were condensed by passing
them through a 5 long stainless steel tube ( I.D) enclosed
by a jacket of coolant (50% ethylene glycol solution
maintained at -8 C) before reaching a set of three glass
impingers kept in the constant temperature bath maintained at
-8 C. Reactor schematic is shown in Figure 1.
The fast pyrolysis of pine wood particles was carried out at
450 oC. The fast pyrolysis vapors were passed through a bed
of ceramic filter (-alumina (30 g) to remove the very fine
char and inorganics before reaching the catalyst bed. The
ceramic filter was followed by a bed of catalyst (20g). Then,
another bed layer of ceramic filter (20 g) was put before the
final bed of catalyst (30g) in the packed bed reactor. All the
fixed bed catalytic experiments had the same amount of
ceramic filter (50 g) and catalyst loading (50 g). A continuous
N2 flow at 1L/min was used to maintain inert atmosphere and
to facilitate bio-oil vapor flow through the catalyst bed. A
blank run, where the PBR was not loaded with any catalyst,
was performed to compare the thermal treatment of pyrolysis
vapors with the catalytic upgrading experiments.
Four different catalysts were screened for their efficacy in
deoxygenating the pyrolysis vapors. A blank experiment
(without catalyst) was performed in order to test the effect of
thermal treatment on deoxygenating the pyrolytic vapors,
whereas the experiment carried out using only the -alumina
support was to evaluate the influence activity of the ceramic
support on bio-oil composition. The screening experiments
were performed at 550 oC at a weight hourly space velocity
(WHSV) of 40 h-1, except in the case of HZSM5 where the
WHSV was kept at 25 h-1.
Characterization of upgraded pyrolysis products

GC/MS analysis
Analysis of aqueous and organic fractions of the liquid
products (bio-oils) was carried out on a Hewlett Packard
5890series II Gas chromatograph/5971 series A mass
spectrometer. The injector temperature was 270 oC. A 30 m x
0.32 mm internal diameter x 0.25 lm thickness silica
capillary column coated with 5% phenylmethylpolysiloxane
was used for the separation of bio-oil compounds. The initial
temperature of the column was set at 40 oC (4 min hold)
followed by heating at 5 oC/min to a nal temperature of 280
o
C. The mass spectrometer employed a 70 eV electron impact
ionization mode, a source (detector) temperature of 250 oC
and an interface temperature of 270 oC.
A 0.1 mg representative samples of bio-oil were diluted with
10 ml of methanol. Only 1 ml of these solutions was then
transferred to the autosampler vial and 10 l of internal
standard (4000 g/ml concentration) was added. Then 2.0 l
of this sample was injected into the GC to acquire the
respective chromatogram. Six isotopically labeled
compounds (US 108N, Ultra Scientic), including 1,
4-dichlorobenzene-d4, naphthalene-d8, acenapthene-d10,
phenanthrene-d10, chrysene-d12 and perylene-d12, were
employed as internal standards to verify the retention times
and quantitate the amounts of thirty known compounds from
a previously published list [40]. From the quantitative data,
the mass (g) of the compound of interest present in 1 g of
liquid product obtained from an experiment is calculated
using the following equation:

Where C is the concentration


of the compound of interest
obtained from the quantitative analysis of the liquid sample
using GC-MS and S is the amount of liquid product sample
dissolved in 10 mL of solvent for GC-MS analysis (g). These
computations were performed for the liquid product from
each catalytic run as well as from the blank run.
Physical analysis
The percent water, acid values, and elemental analysis of the
bio-oil samples were determined. Karl Fisher titration
(Cole-Parmer Model C-25800-10) was performed to measure
the water percentage. The acid values were determined by

36
JOURNAL 0F FOREST PRODUCTS AND INDUSTRIES, 2015, 4(2), 33-43 ISSN: 23254513(print) ISSN 2325 - 453X (Online)

titrating 1 g of the bio-oil samples dissolved in 50/50 v/v


isopropanol/water with 0.1 N NaOH to a pH 8.5. Elemental
analyses (carbon, hydrogen, nitrogen, and oxygen (by
subtraction) of the bio-oil samples contents were measured
with a CE-440 Elemental Analyzer (Exeter Analytical, MA,
USA).

III. RESULTS AND DISCUSSION


Catalytic Effects on the Yields of Pyrolysis Products and the
Physical Properties of the Bio-oils
The effect of catalyst composition on the distribution of
pyrolysis products is shown in Figure 2. Compared to
non-catalytic fast pyrolysis, application of catalysts in fast
pyrolysis increased gas yields at the expense of liquid
products. Among the catalysts, amorphous Si/Al (52%) and
-alumina (47.2%) produced higher liquid yields than the
zeolite catalysts- HZSM5 (41%) and ZY (44.2%). Char yields
(15%) remained constant for all the experiments since char is
a by-product of fast pyrolysis performed in auger reactor and
application of catalysis using a PBR does not influence char
production.
Physical properties of the bio-oils from non-catalytic and
catalytic fast pyrolysis experiments are shown in Table 2.
The water content in the bio-oils varied from 42.8 to 66.4 %.,
with the HZSM5 catalyzed bio-oil having the highest water
content. The raw bio-oil produced in the absence of catalyst
had 48 % water. As upgrading proceeds and more
gasification and deoxygenation occurs, one expects more
water to form. The % water values for the catalyzed bio-oils
indicate that dehydration reactions have occurred.
Dehydration reactions are one of the favorable reaction types
on the surfaces of the acid catalysts possessing bronsted
acidity, resulting in the formation of more water in the
product bio-oil. With regard to acid values, compared to the
raw bio-oil, all catalyzed bio-oils had lower acid values
(when -alumina was used, no change in acid value was
observed). From the acid values of the product bio-oils, it is
easy to notice that HZSM5 was most effective in
decarboxylation and decarbonylation reactions resulting in
the formation of CO, CO2 in the gas phase. Elemental
analysis results show that HZSM5 catalyzed bio-oil
possessed lowest amount of oxygen (30%) while Si/Al
catalyzed bio-oil had highest oxygen content (38%).
Therefore, from the physical properties of catalyzed bio-oils,
HZSM5 was found to be an effective deoxygenation catalyst

for upgrading pine wood fast pyrolysis vapors. However, the


liquid yields were relatively poor and further improvements
are necessary.
Catalytic Effects on the Distribution of Pyrolysis Products
in the Catalyzed Bio-oils
The pyrolysis reactions of the cellulose and hemicellulose
components of biomass produces anhydrosugars including
levoglucosan, levoglucosenone, furans, cyclopentenones,
char, and non-condensable gases, etc. In general, cellulose
pyrolysis was found to follow the Broido- Shafizadeh model,
which involves two competing reaction pathways [44-46].
The first step is the formation of active cellulose which either
transforms to char and non-condensable gases (NCGs) or
depolymerizes to form levoglucosan via transglycolysation
reactions [47-49]. Recently, a third reaction pathway for the
conversion of active cellulose to hydroxyacetaldehyde and
1-hydroxy-2-propanone was also shown [50]. Hemicellulose,
though, assumed to follow cellulose in its pyrolysis
mechanisms [51], due to the structural difference compared to
cellulose, forms furan derivatives and carbonyl compounds
such as cyclopentenones instead of levoglucosan and related
compounds [48, 52-55]. The cellulose and anhydrosugars are
thermally unstable compounds and, on conatct with the
catalyst surface undergoes ring scission reactions to form
furan compounds and decomposition reactions to form char
and non-condensable gases.
No anhydrosugars were noticed in the catalyzed bio-oils.
High catalyst bed temperature (550 oC) employed for
catalytic treatment of fast pyrolysis vapors could be for the
conversion of anhydrosugars to furan compounds, carbon
residue and NCGs. Shen et al. [56] also observed that high
reaction temperatures decrease anhydrosugars yields with
relative increase in the yields of furan compounds and
NCGs. The ring scission reactions of cellulose result in the
formation of 5-hydroxymethyl furfural (HMF) [48], which is
further converted/cracked to 5-methyl furfural, furfuryl
alcohol, furfural [57], furanone, etc. The yields of such
compounds vary depending on the cracking activity of the
catalysts. Greater conversion of anhydrosugars to furfural
derivatives can also be related to the formation of more water
at higher catalyst temperatures; conversion of levoglucosan to
furfural was shown to be facilitated by the presence of water
in the reaction medium [58]. Of the furan compounds,
5-methyl furfural (dehydration product of HMF) was the
most dominant compound in all the experiments except when

38
JOURNAL 0F FOREST PRODUCTS AND INDUSTRIES, 2015, 4(2), 33-43 ISSN: 23254513(print) ISSN 2325 - 453X (Online)

HZSM-5 and -Alumina catalysts were used (Figure 3).


Instead, more furfurals were noticed in ZY and HZSM-5
catalyzed bio-oils, which is an indication of higher cracking
activity of these catalysts. ZY was also active in the formation
of 5-methylfurfural. No specific trends were noticed in the
formation of cyclopentenones and furanones.

experiments. High space velocities (low residence times)


does not allow the fast pyrolysis vapors to spend enough time
on the catalyst reaction sites on the surface or with in the
catalyst pores for deoxygenation to occur.

Figure 4 shows the distribution of pyrolysis products in the


bio-oils produced by non-catalytic and catalytic pyrolysis
experiments. Catechols were the most dominating
compounds in all the catalytic experiments except when Si-Al
was used. Si/Al favored the production of furans and
cyclopentenones while being least effective in converting
heavy compounds such as guaiacols, eugenols, vanillins, and
catechols to low molecular weight phenols. This low cracking
activity of the Si/Al catalyst could also be the reason for high
liquid yields obtained using this catalyst. -alumina produced
the lowest yields of furans and cyclopentenones but formed
highest yields of catechols. The conversion of guaiacol type
heavy compounds to catechols could be the reason for this.
HZSM5 and ZY catalyzed bio-oils also showed low amountss
of guaiacols. But, unlike -alumina, zeolites showed higher
conversion of catechols to phenolics which clearly indicates
theier cracking efficiency.

A larger amount of gas is expected to be produced during the


combined auger reactor pyrolysis plus vapor phase catalytic
treatment because the catalytic reactor (550 oC) subjected the
vapors to temperatures higher than the 450 C fast wood
pyrolysis temperature, thereby further gasifying some of the
products. While the entire gaseous product was not collected,
sample bags were used to periodically effluent gas during the
run with HZSM5 catalyst. Two gas samples were collected
and analyzed using GC equipped with TCD and FID
detectors. The product distribution of two gas samples
collected during the run is shown in Table 3. The gas
compositions accounted to 87% and 78% of the total gases of
sample 1 and sample 2, respectively. The rest of the gases,
assumed to be C2+ gaseous hydrocarbons, are not reported
because the GC was not calibrated for their quantitative
analysis. CO is the dominant component of the gas
compositions followed by CO2, indicating that main routes
for deoxygenation of pyrolytic vapors occurs through
decarbonylation and decarboxylation reactions.

HZSM5 catalyzed bio-oil had more phenols among all the


bio-oils from catalytic experiments. Cracking of guaiacols,
eugenols and vanillins to catechols, which are further cracked
to phenols is the predominant reaction pathway that leads to
the formation of phenolics during catalytic vapor upgrading
reactions. Phenols were earlier reported [21] to be the
primary reaction products in cracking of pyrolysis oil over
acidic HZSM5 catalyst, but the present results were found to
differ, where diols (catechols) were found to be predominant
compared to phenols. Use of much lower WHSVs (3.75 h-1)
in those studies [21] could be the reason for such variation in
the chemical composition of upgraded bio-oils. Regarding the
non-oxygenated aromatic hydrocarbons, HZSM5 was the
only catalyst that showed formation of toluene, xylene,
indene, as well as polyaromatic hydrocarbons such as
naphthalene and 2-methyl naphthalene. The formation of
such aromatics is believed to be through a carbenium ion
mechanism, i.e. by oligomerization of cracked fragments,
alkylation, ismomerization, cyclization, followed by
aromatization. However, the yields (GC peak area - 7%) of
non-oxygenated hydrocarbons were poor. Such poor yields
of hydrocarbons might be due to the high WHSVs (40 h-1)
and high catalyst bed temperatures (550 oC) employed in our

Gas Analysis

The catalyst screening experiments resulted in mild


deoxygenation of pyrolytic vapors through dehydration,
decarboxylation, and decarbonylation reactions. Of the four
catalysts, -alumina and HZSM5 catalysts perfomed well in
decreasing the carbonyl function groups and converting the
heavy oxygenates like guaiacols to either catechols or
phenols. Carbonyl groups (aldehydes, ketones) are one of the
very reactive functional groups present in the bio-oil and their
elimination increases thermal/storage stability of the
upgraded bio-oil. HZSM5 was most effective in decreasing
the acid value as well as obtaining value added chemicals like
phenols. In addition, HZSM5 also favored the formation of
hydrocarbons, though the yields were poor. From these
results, it can be inferred that HZSM5 is an effective cracking
catalyst. Since the deoxygenation reactions and the
production of non-oxygented hydrocarbons was poor, further
investigation on the efficacy of HZSM5 catalyst was carried
out by varying the process variables like WHSV and reaction
temperture.
Thermogravimetric Analysis (TGA) of Used Catalysts

38
JOURNAL 0F FOREST PRODUCTS AND INDUSTRIES, 2015, 4(2), 33-43 ISSN: 23254513(print) ISSN 2325 - 453X (Online)

TGA was perfomed on the used catalysts to determine the


amount of residual carbon deposited on the catalyst surface.
A Shimadzu instrument TGA-50 was used to perform the TG
analysis. A required amount of catalyst was placed on an
alumina pan and a temperature program form room
temperature to 900 oC at a ramping rate of 5 oC/min was used.
The runs were performed under a N2 flow of 30 mL/min.
Percentage weight losses of the catalysts used in pyrolysis
vapor upgrading experiments are shown below in Figure 5.
From Figure 5, it can be seen that the carbon deposition was
higher with crystalline HZSM5 and ZY catalysts compared to
Si/Al and -alumina catalysts. Ordered pore structures, in
addition to pore specific and shape selective catalysis, of
zeolite catalysts could be attributed for their high coking
activity. The coke components are usually polyaromatics at
high temperatures (>350 oC) and their formation involves
hydrogen transfer (acid catalysts) and dehydrogenation
(bifunctional catalysts) steps in addition to condensation and
rearrangement steps [59]. The most significant step in coking
of zeolites, especially large pored zeolites such as ZY, is
assumed to be the alkylation of aromatics [60-63]. On the
active reaction sites (either catalyst surface or within the
pores) of HZSM5, the alkyl aromatics undergo cyclization
reactions into fused-ring polycyclics, which eventually
dehydrogenate to coke [64]. TGA results show that the
catalysts can be completely regenerated at 650 oC

Low liquid yields and high coking of zeolites are the


drawbacks involved in the current study. Significant further
improvements are possible by optimizing the process
variables, which will be the focus of our future studies.

ACKNOWLEDGEMENT
This research was supported by the Sustainable Energy
Research Center at Mississippi State University and the U.S.
Department of Energy through DOE-DE-FG36-06GO86025.

DISCLAIMER
This report was prepared as an account of work sponsored by
an agency of the United States government. Neither the
United States Government nor any agency thereof, nor any of
their employees, makes any warranty, express or implied, or
assumes any legal liability or responsibility for the accuracy,
completeness, or usefulness of any information, apparatus,
product, or process disclosed, or represents that its use would
not infringe privately owned rights. Reference herein to any
specific commercial product, process, or service by trade
name, trademark, manufacturer, or otherwise does not
necessarily constitute or imply its endorsement,
recommendation, or favoring by the United States
Government or any agency thereof. The views and opinions
of authors expressed herein do not necessarily state or reflect
those of the United States Government or any agency thereof.

IV. CONCLUSIONS
The effects of acid catalysts were tested in upgrading pine
wood fast pyrolysis vapors using an auger-packed bed
integrated reactor system. Auger reactor was used for fast
pyrolysis of pine wood and packed bed reactor was used for
catalytic treatment of pyrolysis vapors coming out of auger
reactor. Among Si/Al, -Alumina, HZSM5 and ZY catalysts,
HZSM5 was a superior deoxygenatin catalyst with its
upgraded bio-oil having lowest acid number (46.4%), lowest
oxygen content (30%) and high amount of phenolics among
all the catalyzed bio-oils. Aromatic hydrocarbons, though
formed in very low yields (7% of total GC relative peak area),
was seen only in HZSM5 catalyzed bio-oil. However, the
liquid yields were the lowest (42%) using HZSM5 catalyst
while Si/Al produced the highest yields (52%). TGA
indicated that zeolites coked more compared to amorphous
Si/Al and alumina catalysts; large pored ZY showed highest
coking activity. The catalysts can be completely regenerated
by calcining them at 650 oC.

REFERENCES
[1.]

[2.]

[3.]

[4.]

[5.]

[6.]

[7.]

Bridgwater AV. Renewable fuels and chemicals by thermal


processing of biomass. Chemical Engineering Journal
2003;91(2-3):87-102.
Bulushev DAand Ross JRH. Catalysis for conversion of biomass
to fuels via pyrolysis and gasification: A review. Catalysis Today
2011;171(1):1-13.
Guda VK, Steele PH, Penmetsa VKand Li Q. Chapter 7 - Fast
Pyrolysis of Biomass: Recent Advances in Fast Pyrolysis
Technology. In: Sukumaran APBSK, editor. Recent Advances in
Thermo-Chemical Conversion of Biomass. Boston: Elsevier;
2015. p. 177-211.
Zhang Q, Chang J, Wang Tand Xu Y. Review of biomass
pyrolysis oil properties and upgrading research. Energy
Conversion and Management 2007;48(1):87-92.
Lu Q, Li W-Zand Zhu X-F. Overview of fuel properties of
biomass fast pyrolysis oils. Energy Conversion and Management
2009;50(5):1376-1383.
Bridgwater AV. Production of high grade fuels and chemicals
from catalytic pyrolysis of biomass. Catalysis Today
1996;29(1-4):285-295.
Chheda JN, Huber GWand Dumesic JA. Liquid-Phase Catalytic
Processing of Biomass-Derived Oxygenated Hydrocarbons to
Fuels and Chemicals. Angewandte Chemie International Edition
2007;46(38):7164-7183.

39
JOURNAL 0F FOREST PRODUCTS AND INDUSTRIES, 2015, 4(2), 33-43 ISSN: 23254513(print) ISSN 2325 - 453X (Online)

[8.]

[9.]

[10.]

[11.]

[12.]

[13.]

[14.]

[15.]
[16.]

[17.]

[18.]

[19.]

[20.]

[21.]

[22.]

[23.]

[24.]

[25.]

[26.]

[42.]

Li H, Xu Q, Xue Hand Yan Y. Catalytic reforming of the aqueous


phase derived from fast-pyrolysis of biomass. Renewable Energy
2009;34(12):2872-2877.
Vispute TPand Huber GW. Production of hydrogen, alkanes and
polyols by aqueous phase processing of wood-derived pyrolysis
oils. Green Chemistry 2009;11(9):1433-1445.
Huber GWand Dumesic JA. An overview of aqueous-phase
catalytic processes for production of hydrogen and alkanes in a
biorefinery. Catalysis Today 2006;111(1-2):119-132.
Piskorz J, Radlein DSAG, Scott DSand Czernik S. Pretreatment
of wood and cellulose for production of sugars by fast pyrolysis.
Journal of Analytical and Applied Pyrolysis 1989;16(2):127-142.
Huber GW, Chheda JN, Barrett CJand Dumesic JA. Production of
Liquid
Alkanes
by
Aqueous-Phase
Processing
of
Biomass-Derived
Carbohydrates.
Science
2005;308(5727):1446-1450.
Gong F, Yang Z, Hong C, Huang W, Ning S, Zhang Z, Xu Yand
Li Q. Selective conversion of bio-oil to light olefins: Controlling
catalytic cracking for maximum olefins. Bioresource Technology
2011;102(19):9247-54.
Carlson TR, Tompsett GA, Conner WCand Huber GW. Aromatic
Production from Catalytic Fast Pyrolysis of Biomass-Derived
Feedstocks. Topics in Catalysis 2009;52(3):241-252.
Elliott DC. Historical Developments in Hydroprocessing
Bio-oils. Energy & Fuels 2007;21(3):1792-1815.
Parapati DR, Guda VK, Penmetsa VK, Steele PHand Tanneru
SK. Single stage hydroprocessing of pyrolysis oil in a continuous
packed-bed reactor. Environmental Progress & Sustainable
Energy 2014;33(3):726-731.
Parapati DR, Guda VK, Penmetsa VK, Tanneru SK, Mitchell
Band Steele PH. Comparison of reduced and sulfided
CoMo/-Al2O3 catalyst on hydroprocessing of pretreated bio-oil
in a continuous packed-bed reactor. Environmental Progress &
Sustainable Energy 2014:n/a-n/a.
Park HJ, Jeon J-K, Suh DJ, Suh Y-W, Heo HSand Park Y-K.
Catalytic Vapor Cracking for Improvement of Bio-Oil Quality.
Catalysis Surveys from Asia 2011;15(3):161-180.
Adjaye JDand Bakhshi NN. Production of hydrocarbons by
catalytic upgrading of a fast pyrolysis bio-oil. Part I: Conversion
over various catalysts. Fuel Processing Technology
1995;45(3):161-183.
Adjaye JD, Katikaneni SPRand Bakhshi NN. Catalytic
conversion of a biofuel to hydrocarbons: effect of mixtures of
HZSM-5 and silica-alumina catalysts on product distribution.
Fuel Processing Technology 1996;48(2):115-143.
Samolada MC, Papafotica Aand Vasalos IA. Catalyst Evaluation
for Catalytic Biomass Pyrolysis. Energy & Fuels
2000;14(6):1161-1167.
Liu C, Wang H, Karim AM, Sun Jand Wang Y. Catalytic fast
pyrolysis of lignocellulosic biomass. Chemical Society Reviews
2014.
Guda VKand Penmetsa VK. Chapter 1 -Catalytic Fast Pyrolysis
of Lignocellulosic Biomass: Review and Recent Progress In:
Ambrosio J, editor. Handbook on Oil Production Research: Nova
Science Publishers; 2014. p. 1-38.
Srinivas ST, Dalai AKand Bakhshi NN. Thermal and catalytic
upgrading of a biomass-derived oil in a dual reaction system. The
Canadian Journal of Chemical Engineering 2000;78(2):343-354.
Carlson TR, Vispute TPand Huber GW. Green gasoline by
catalytic fast pyrolysis of solid biomass derived compounds.
ChemSusChem 2008;1(5):397-400.
Antonakou E, Lappas A, Nilsen MH, Bouzga Aand Stcker M.
Evaluation of various types of Al-MCM-41 materials as catalysts
in biomass pyrolysis for the production of bio-fuels and
chemicals. Fuel 2006;85(14-15):2202-2212.
Li B, Lv W, Zhang Q, Wang Tand Ma L. Pyrolysis and catalytic
upgrading of pine wood in a combination of auger reactor and
fixed bed. Fuel 2014;129(0):61-67.

[27.]

[28.]

[29.]

[30.]

[31.]

[32.]

[33.]

[34.]

[35.]

[36.]

[37.]

[38.]

[39.]

[40.]

[41.]

[43.]

Iliopoulou EF, Antonakou EV, Karakoulia SA, Vasalos IA,


Lappas AAand Triantafyllidis KS. Catalytic conversion of
biomass pyrolysis products by mesoporous materials: Effect of
steam stability and acidity of Al-MCM-41 catalysts. Chemical
Engineering Journal 2007;134(1-3):51-57.
Nilsen MH, Antonakou E, Bouzga A, Lappas A, Mathisen Kand
Stcker M. Investigation of the effect of metal sites in
Me-Al-MCM-41 (Me = Fe, Cu or Zn) on the catalytic behavior
during the pyrolysis of wooden based biomass. Microporous and
Mesoporous Materials 2007;105(1-2):189-203.
Triantafyllidis KS, Iliopoulou EF, Antonakou EV, Lappas AA,
Wang Hand Pinnavaia TJ. Hydrothermally stable mesoporous
aluminosilicates (MSU-S) assembled from zeolite seeds as
catalysts for biomass pyrolysis. Microporous and Mesoporous
Materials 2007;99(1-2):132-139.
Heo HS, Kim SG, Jeong KE, Jeon JK, Park SH, Kim JM, Kim
SSand Park YK. Catalytic upgrading of oil fractions separated
from food waste leachate. Bioresource Technology
2011;102(4):3952-7.
Adam J, Blazs M, Mszros E, Stcker M, Nilsen MH, Bouzga
A, Hustad JE, Grnli Mand ye G. Pyrolysis of biomass in the
presence
of
Al-MCM-41
type
catalysts.
Fuel
2005;84(12-13):1494-1502.
Adam J, Antonakou E, Lappas A, Stcker M, Nilsen MH, Bouzga
A, Hustad JEand ye G. In situ catalytic upgrading of biomass
derived fast pyrolysis vapours in a fixed bed reactor using
mesoporous materials. Microporous and Mesoporous Materials
2006;96(1-3):93-101.
Park HJ, Heo HS, Jeon J-K, Kim J, Ryoo R, Jeong K-Eand Park
Y-K. Highly valuable chemicals production from catalytic
upgrading of radiata pine sawdust-derived pyrolytic vapors over
mesoporous MFI zeolites. Applied Catalysis B: Environmental
2010;95(3-4):365-373.
Park K-H, Park HJ, Kim J, Ryoo R, Jeon J-K, Park Jand Park
Y-K. Application of Hierarchical MFI Zeolite for the Catalytic
Pyrolysis of Japanese Larch. Journal of Nanoscience and
Nanotechnology 2010;10(1):355-359.
Jae J, Coolman R, Mountziaris TJand Huber GW. Catalytic fast
pyrolysis of lignocellulosic biomass in a process development
unit with continual catalyst addition and removal. Chemical
Engineering Science 2014;108:33-46.
Zhang H, Xiao R, Jin B, Shen D, Chen Rand Xiao G. Catalytic
fast pyrolysis of straw biomass in an internally interconnected
fluidized bed to produce aromatics and olefins: Effect of different
catalysts. Bioresource Technology 2013;137(0):82-87.
Zhang H, Xiao R, Jin B, Xiao Gand Chen R. Biomass catalytic
pyrolysis to produce olefins and aromatics with a physically
mixed catalyst. Bioresource Technology 2013;140(0):256-262.
Zhang H, Carlson TR, Xiao Rand Huber GW. Catalytic fast
pyrolysis of wood and alcohol mixtures in a fluidized bed reactor.
Green Chemistry 2012;14(1):98.
Pittman CU, Mohan D, Eseyin A, Li Q, Ingram L, Hassan E-BM,
Mitchell B, Guo Hand Steele PH. Characterization of Bio-oils
Produced from Fast Pyrolysis of Corn Stalks in an Auger Reactor.
Energy & Fuels 2012;26(6):3816-3825.
Ingram L, Mohan D, Bricka M, Steele P, Strobel D, Crocker D,
Mitchell B, Mohammad J, Cantrell Kand Pittman CU. Pyrolysis
of Wood and Bark in an Auger Reactor: Physical Properties and
Chemical Analysis of the Produced Bio-oils. Energy & Fuels
2007;22(1):614-625.
Thangalazhy-Gopakumar S, Adhikari S, Gupta RBand Fernando
SD. Influence of Pyrolysis Operating Conditions on Bio-Oil
Components: A Microscale Study in a Pyroprobe. Energy &
Fuels 2011;25(3):1191-1199.
Veses A, Aznar M, Martnez I, Martnez JD, Lpez JM, Navarro
MV, Calln MS, Murillo Rand Garca T. Catalytic pyrolysis of

40
JOURNAL 0F FOREST PRODUCTS AND INDUSTRIES, 2015, 4(2), 33-43 ISSN: 23254513(print) ISSN 2325 - 453X (Online)

[44.]

[45.]

[46.]

[47.]

[48.]
[49.]

[50.]

[51.]
[52.]

[53.]

wood biomass in an auger reactor using calcium-based catalysts.


Bioresource Technology 2014;162(0):250-258.
Broido A, Evett Mand Hodges CC. Yield of
1,6-anhydro-3,4-dideoxy--d-glycero-hex-3-enopyranos-2-ulose
(levoglucosenone) on the acid-catalyzed pyrolysis of cellulose
and 1,6-anhydro--d-glucopyranose (levoglucosan). Carbohydr
Res 1975;44(2):267-274.
Shafizadeh Fand Bradbury AGW. Thermal degradation of
cellulose in air and nitrogen at low temperatures. Journal of
Applied Polymer Science 1979;23(5):1431-1442.
Bradbury AGW, Sakai Yand Shafizadeh F. A kinetic model for
pyrolysis of cellulose. Journal of Applied Polymer Science
1979;23(11):3271-3280.
Arseneau DF. Competitive Reactions in the Thermal
Decomposition of Cellulose. Canadian Journal of Chemistry
1971;49(4):632-638.
Akio Ohnishi KK, and Eriko Takagi. Curie-Point Pyrolysis of
Cellulose. Polymer Journal 1975;7(4):431-437.
Milosavljevic I, Oja Vand Suuberg EM. Thermal Effects in
Cellulose Pyrolysis: Relationship to Char Formation Processes.
Industrial
&
Engineering
Chemistry
Research
1996;35(3):653-662.
Luo, Wang, Liaoand Cen. Mechanism Study of Cellulose Rapid
Pyrolysis. Industrial & Engineering Chemistry Research
2004;43(18):5605-5610.
Ponder GRand Richards GN. Thermal synthesis and pyrolysis of
a xylan. Carbohydr Res 1991;218(0):143-155.
Zhu X, Lu Q, Li Wand Zhang D. Fast and catalytic pyrolysis of
xylan: Effects of temperature and M/HZSM-5 (M = Fe, Zn)
catalysts on pyrolytic products. Frontiers of Energy and Power
Engineering in China 2010;4(3):424-429.
Shen DK, Gu Sand Bridgwater AV. Study on the pyrolytic
behaviour of xylan-based hemicellulose using TGFTIR and

Figure 1. Schematic of auger packed bed integrated reactor


system used for catalytic vapor upgrading experiment

[54.]

[55.]

[56.]

[57.]

[58.]

[59.]
[60.]
[61.]

[62.]

[63.]
[64.]

PyGCFTIR. Journal of Analytical and Applied Pyrolysis


2010;87(2):199-206.
Guo X, Wang S, Zhou Yand Luo Z. Catalytic pyrolysis of
xylan-based hemicellulose over zeolites. INTERNATIONAL
JOURNAL of ENERGY and ENVIRONMENT 2011;5(4):524-531.
Gao-Jin L, Shu-Bin Wand Rui L. Characteristics of corn stalk
hemicellulose pyrolysis in a tubular reactor. BioResources
2010;5(4):2051-2062.
Shen DKand Gu S. The mechanism for thermal decomposition of
cellulose and its main products. Bioresource Technology
2009;100(24):6496-504.
Shafizadeh Fand Lai YZ. Thermal degradation of
1,6-anhydro-.beta.-D-glucopyranose. The Journal of Organic
Chemistry 1972;37(2):278-284.
Kawamoto H, Saito S, Hatanaka Wand Saka S. Catalytic
pyrolysis of cellulose in sulfolane with some acidic catalysts.
Journal of Wood Science 2007;53(2):127-133.
Guisnet Mand Magnoux P. Organic chemistry of coke formation.
Applied Catalysis A: General 2001;212(1-2):83-96.
Walsh DEand Rollmann LD. Radiotracer experiments on carbon
formation in zeolites. Journal of Catalysis 1977;49(3):369-375.
Walsh DEand Rollmann LD. Radiotracer experiments on carbon
formation
in
zeolites.
II.
Journal
of
Catalysis
1979;56(2):195-197.
Bibby DM, Milestone NB, Patterson JEand Aldridge LP. Coke
formation in zeolite ZSM-5. Journal of Catalysis
1986;97(2):493-502.
Bhatia S, Beltramini Jand Do DD. Deactivation of Zeolite
Catalysts. Catalysis Reviews 1989;31(4):431-480.
Sigmund M C. Shape-selective catalysis in zeolites. Zeolites
1984;4(3):202-213.

41

I-1
I-5 I-9
I-10

I-3
I-4
I-12
I-2

Exit non-condensable
gases

Pump

Condenser

I-7
I-6 I-13

Biomass
hopper

I-8
I-11

N2

Fixed-bed
catalytic reactor

JOURNAL 0F FOREST PRODUCTS AND INDUSTRIES, 2015, 4(2), 33-43 ISSN: 23254513(print) ISSN 2325 - 453X (Online)

Gas
sampling

Auger reactor
for
fast pyrolysis

Bio-oil

Figure 2. Effect of catalyst compoition on the yields of pyrolysis products

Figure 3. Distribution of furans and cyclopentenones in


the bio-oils obtained from catalytic vapor upgrading experiment

42
JOURNAL 0F FOREST PRODUCTS AND INDUSTRIES, 2015, 4(2), 33-43 ISSN: 23254513(print) ISSN 2325 - 453X (Online)

Figure 4. Distribution of pyrolysis products in the bio-oils obtained from catalytic vapor upgrading experiments

Figure 5. TGA of the catalysts used in catalytic vapor upgrading experiment

Table 1: Catalysts used in the catalytic upgrading of pine wood fast pyrolysis vapors.
Catalyst

SiO2/Al2O3 ratio

Surface
Area (m2/g)

Average Pore
Size (nm)

Structure

CBV2314ZSM5

23

425

0.50 0.55

3-D pore system; Cubic crystal structure with straight


10-membered ring channels (5.3 x 5.5 Ao) connected by
sinusoidal channels (5.1 x 5.5Ao)

CBV720Zeolite Y

30

780

0.70

3-D pore system; Cubic crystal structure with 12-membered ring


channels (7.4 Ao) connecting spherical 11.8 Ao cavities
(supercages)

Si/Al

5.15

436

-alumina

99% alumina

245

43
JOURNAL 0F FOREST PRODUCTS AND INDUSTRIES, 2015, 4(2), 33-43 ISSN: 23254513(print) ISSN 2325 - 453X (Online)

Table 3: Product distribution of gas samples


(Catalyst: HZSM5; Temperature: 550 C; WHSV: 25 h-1).
Gas

Sample-1

Sample-2

Hydrogen (%)

8.4

7.0

Carbon monoxide (%)

40.8

27.7

Carbon dioxide (%)

15.4

12

Methane (%)

9.8

6.4

Nitrogen (%)

13

25

Table 2: Physical properties of the bio-oils obtained from


in-situ catalytic treatment of pine wood pyrolysis vapors
Elemental Composition

Catalyst

Acid Value
(mg KOH/g
Oil)

Water
(%)

Untreated

81

20

-alumina

82.1

56.8

Blank run

77.8

48

Si/Al

74.1

-Alumina

C
%

H
%

N
%

O%

42.8

53.9

7.5

0.03

38.5

61.2

60.4

55.5

7.1

0.6

36.8

ZY

61.0

59.6

54.2

7.4

0.8

37.6

HZSM5

46.4

66.4

61.1

7.9

0.5

30.4

Das könnte Ihnen auch gefallen