Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

DOI: 10.1002/cctc.

201500840 Full Papers

Simultaneous Liquefaction and Hydrodeoxygenation of


Lignocellulosic Biomass over NiMo/Al2O3, Pd/Al2O3, and
Zeolite Y Catalysts in Hydrogen Donor Solvents
Miha Grilc, Blaž Likozar,* and Janez Levec[a]

The one-step solvolysis and hydro-treatment of oak, fir and sis and an innovative method for particle size distribution and
beech sawdust was studied in a slurry reactor using tetralin, mean grain size determination by SEM image processing. A
phenol and glycerol as solvents and representative heteroge- new lumped kinetic model was developed according to the re-
neous catalysts used in the petrochemical industry for hydro- action mechanisms, which contain wood de-polymerisation,
genolysis (the sulfide form of NiMo/Al2O3), hydrogenation (Pd/ decarbonylation, decarboxylation and demethanation, tar
Al2O3) or fluid catalytic cracking (zeolite Y). Deoxyliquefaction phase hydrodeoxygenation by molecular and in situ generated
products of cellulose, hemicellulose and lignin were character- hydrogen, as well as charring inhibition by free-radical stabili-
ised in terms of solvent fractionation, FTIR and diffuse reflec- sation hydrogen transfer.
tance infrared Fourier transform spectroscopy, elemental analy-

Introduction

The valorisation of waste lignocellulosic (LC) biomass by its catalyst with an active phase (e.g., MoS2, WS2, Ni, Ru, Pt or Pd)
conversion into higher-value chemicals or liquid transportation that is responsible for HDO and hydrogenation, whereas an
fuels has attracted great attention in recent years. A typical acidic support (Al2O3, SiO2, etc.) is responsible for bond cleav-
thermochemical approach to convert the LC biomass into age, required for heteroatom removal and catalytic cracking re-
liquid transportation fuel involves a two-step process, in which actions.[4] The addition of a hydrogen donor solvent is also de-
solid biomass or its components are de-polymerised in the first sirable to improve hydrogen transfer and stabilise free radicals
step by pyrolysis, dissolution/liquefaction (LIQ) or gasification, formed during catalytic cracking.[4a, 5] The similarity of the reac-
and low-value intermediates (i.e., pyrolysis oil, liquefaction oil tion conditions for direct liquefaction and HDO, which include
or synthesis gas) are subsequently upgraded by catalytic the reaction temperatures in the range of 250–400 8C, elevated
hydro-treatment or Fischer–Tropsch synthesis to obtain a com- hydrogen pressure and a beneficial influence of hydrogen
petitive-quality liquid fuel.[1] In contrast to pyrolysis or gasifica- donor solvent, makes the integration of liquefaction processes
tion that both require high operating temperatures, direct liq- possible and presumably feasible. Rocha et al.[6] investigated
uefaction is performed at a significantly lower temperature in the solvent-free pyrolysis and HDO of cellulose under an ele-
the presence of solvents that may serve as reactants and heat vated H2 pressure in the single- and two-stage process by
carriers, whereas the use of hydrogen donor solvents, accom- using dispersed iron sulfide or the sulfide form of NiMo/Al2O3
panied by a moderate H2 pressure, is known to promote C¢C and reported an increased semi-solid tar yield from 8 to
and C¢O bond cleavage, to reduce coke formation and inhibit 32 wt % by increasing the reaction temperature from 300 to
re-polymerisation reactions.[2] A higher temperature is expect- 350 8C. A further temperature increase to 520 8C only affected
ed to be required for the biomass liquefaction using tetralin in the char and gas yield, and the tar fraction remained more or
comparison to phenol, glycerol or other polyhydroxyl alcohols, less unchanged.[6] Vasilakos and Austgen[2a] also investigated si-
therefore, liquefaction conditions might already overlap with multaneous cellulose liquefaction and HDO under H2 or an
pyrolysis, whereas the absence of the chemically bonded inert atmosphere, and tetralin was used as a solvent. Oil-phase
oxygen in this solvent avoids its removal in the hydrodeoxyge- yields as high as 52 wt % were reported without the use of
nation (HDO) step.[3] However, the hydro-treatment of bio- a catalyst, and the addition of Pd/Al2O3 increased the yield.
mass-derived oils also requires temperatures above 300 8C, Little solid residue indicates that cellulose might not be re-
high hydrogen pressure and the presence of a heterogeneous sponsible for potential char formation during the deoxylique-
faction of wood.[2a] The influence of temperature, pressure and
[a] Dr. M. Grilc, Assist. Prof. Dr. B. Likozar, Prof. Dr. J. Levec reaction time on the direct liquefaction of sawdust in tetralin
Laboratory of Catalysis and Chemical Reaction Engineering at an elevated hydrogen pressure but without a catalyst was
National Institute of Chemistry
investigated by Wang et al. and Yan et al. in the temperature
Hajdrihova 19, SI-1001 Ljubljana (Slovenia)
E-mail: blaz.likozar@ki.si range of 150–450 and 200–350 8C, respectively.[7] Araya et al. in-
Supporting Information for this article is available on the WWW under vestigated the direct liquefaction of pine sawdust, cellulose
http://dx.doi.org/10.1002/cctc.201500840. and lignin in tetralin using CoMo/Al2O3, Pt/Al2O3 and Raney Ni

ChemCatChem 2016, 8, 180 – 191 180 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
catalysts and described the influence of feedstock, catalyst According to the SEM image (Figure S1), zeolite Y comprises
type, solvent and atmosphere.[8] However, no reaction mecha- uniform octahedral crystals with an edge length of 600–
nism or kinetic model was proposed as no temperature- or 800 nm. It exhibits a superior specific area because of its
time-dependent results were obtained. However, Beuchet et al. micro-porosity and relatively high pore volume. However,
used a mixture of lignocellulosic wastes and Raney Ni to inves- a high specific surface area determined by BET might not cor-
tigate hydroliquefaction in tetralin under H2 pressure and pro- respond to the actual surface area available for reactions as
posed a schematic pathway for hydroliquefaction that involves pores with a diameter below 10 æ might not be accessible for
two steps: non-catalytic solvolysis by thermal degradation of the rather large molecules present in liquefied biomass. More-
biomass as the first, and the catalytic hydrogenation and HDO over, micro-porous material might also be more sensitive to
enhanced by hydrogen transfer as the second step.[9] physical deactivation because of pore plugging, which is quite
In this work, the simultaneous biomass liquefaction and possible if a slurry bulk that contains the biomass particles and
hydro-treatment of LC biomass was investigated in hydrogen real biomass-derived oils is processed. The actual reactive sur-
donor solvents under an inert or H2 atmosphere in the pres- face area during the process is, therefore, expected to be
ence of a typical HDO, hydrogenation or fluid catalytic cracking lower than that determined by the N2 physisorption method.
(FCC) heterogeneous catalyst. The influence of the reaction
time, temperature, pressure, process gas, catalyst, wood and
Biomass characterisation
solvent type on the yield and composition of products was es-
tablished and a new lumped kinetic model that considers LC The chemical and morphological characteristics of the dried
biomass liquefaction, decarboxylation, decarbonylation, HDO sawdust samples used in this study are presented in Table 2.
and char formation was developed based on the reaction
mechanisms proposed from the experimental results.
Table 2. Properties of biomass feedstock before LIQ-HT

Sample Elemental analysis GCV[b] d50[c] SBET[d]


Results and Discussion [wt %] [MJ kg¢1] [mm] [m2 g¢1]
C H S O[a]
Catalyst characterisation
[e]
Beech 48.1 6.4 0.16 45.3 17.9 31, 191 0.54, 0.28
The characteristics of the pre-treated catalysts, tested in this Fir 47.4 6.0 < LOD[f] 46.6 16.9 173 0.21
work, are summarised in Table 1. NiMo/Al2O3 is a typical bi- Oak 49.1 6.2 < LOD[f] 44.7 18.1 180 0.31
functional catalyst with heteroatom-removal and hydrocrack- [a] Calculated as balance. [b] Gross calorific value calculated using the
ing functionality. The tested catalyst type and off-site sulfide Dulong formula. [c] According to the particle size distribution determined
form pre-rendering method are the same as those used in the by SEM. [d] BET surface area. [e] Two particle size fractions of beech wood
were tested. [f] Limit of detection.
petrochemical industry, and hence, such a catalyst, with equal
performance and physical properties, which include metal
loading and specific surface area, is potentially available (and
feasible) for use on a larger scale. Pd/Al2O3 was chosen as a dis- Elemental analysis showed that all sawdust samples contained
tinct representative of hydrogenation catalysts in the petrole- relatively similar C, H and O amounts and possessed a compara-
um industry. NiMo/Al2O3 and Pd/Al2O3 are both alumina-sup- ble calorific value. The particle size distribution of the sawdust
ported meso-porous materials (the morphology of both mate- samples used in this work is shown in Figure 1, and the SEM
rials is depicted in Figure S1) with similar pore volumes, where- images of some representative sawdust particles are shown in
as the average pore diameter of the latter is somewhat higher. Figure S2. According to the analysis of 1039 fine beech saw-
This might be, along with the larger particle size, the main dust particles, the sample only contained biomass pieces (with
reason for the significantly lower specific surface area of Pd/ the minimal Feret diameter) smaller than 75 mm, whereas the
Al2O3 than that of NiMo/Al2O3. Zeolite Y represents the group other sawdust sample dimensions were significantly larger. To
of FCC catalysts. As it was used in a powder form, the SEM reduce the influence of the eventual mass transfer limitations
technique was used to approximate the crystal size. within the pores of the particles during liquefaction, most of

Table 1. Catalysts used for simultaneous liquefaction and hydro-treatment.

Catalyst Form[a] Pre-treatment Active phase SBET[b] VP[c] DP[d]


[wt %] [m2 g¢1] [cm3 g¢1] [æ]
NiMo/Al2O3 1 mm sulfide formation > 4.5 (NiS), > 22 (MoS2) > 200 > 0.40 < 100
Q
Pd/Al2O3 2–4 mm reduction 0.3 (Pd) 118 0.48 160
S
Zeolite Y 0.6–0.8 mm calcination 73 (SiO2), 24 (Al2O3) 549 0.26 9.4
P

[a] Q = quadri-lobe particles; S = spheres; P = powder. [b] BET surface area. [c] Pore volume. [d] Pore diameter.

ChemCatChem 2016, 8, 180 – 191 www.chemcatchem.org 181 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

Figure 1. Particle size distribution of a) fine beech, b) coarse beech, c) fir, and d) oak sawdust.

the experiments were performed by using the fine beech


wood powder with a median grain size (d50) of only 33 mm.
A small particle size and correspondingly high reactive sur-
face area is expected to have a crucial effect on the solid–
liquid reactions of liquefaction. For this reason the influence of
the wood type on the progress of the reactions involved in
the simultaneous liquefaction and hydro-treatment (LIQ-HT)
was investigated using the beech, fir and oak samples with
comparable particle size distributions and similar specific sur-
face areas.

Effect of the processing time


The influence of the process time on the reactions involved in
Figure 2. Experimental (symbols) and modelling (lines) results for the simul-
LIQ-HT was investigated by conducting the experiments with taneous direct liquefaction and hydrodeoxygenation of beech sawdust over
the fine beech sawdust in tetralin for 30, 60 and 90 min at the the sulfide form of NiMo/Al2O3 at 350 8C, heating rate of 7.5 K min¢1, total
temperature of 350 8C, the hydrogen pressure of 5 MPa and pressure of 5 MPa, stirring speed of 500 rpm, sawdust-to-tetralin mass ratio
of 1:4 and catalyst-to-sawdust mass ratio of 0.10.
using the NiMo/Al2O3 catalyst. Biomass was to a great extent
already liquefied within the first 30 min at the plateau temper-
ature (Tp) to result in the formation of the polar tar phase in
the first step and its subsequent HDO in the second (Figure 2). Gas formation (Figure 3) can be attributed to the biomass
The amount of insoluble residue decreased to 21 wt % at de-polymerisation by either direct liquefaction or pyrolysis as
75 min (30 min after TP was reached) but it apparently con- the rate of (CO2, CO and CH4) gas formation showed a straight-
tained some char as the resulting solid powder was completely forward proportional dependency on the time profile of the
black. The char content increased during the process time as unconverted wood content. CO and CO2 could also have been
the solid residue content reached its minimal value of formed from the oxygen-containing tar phase, but in that case,
0.17 wt % after … 60 min at Tp, whereas a longer hydro-process- their evolution rates would not be completely diminished after
ing time caused its substantial increase. An increase of insolu- 100 min when the tar phase content is still relatively high. As
ble residue content corresponded to the decrease of liquid kinetic modelling could not have provided a satisfactory agree-
HDO products, which indicates that the hydrodeoxygenated ment between the calculated and measured values, if we ac-
oil phase is the predominant precursor for char formation. knowledge the latter reaction mechanism, it was concluded

ChemCatChem 2016, 8, 180 – 191 www.chemcatchem.org 182 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

Figure 3. Experimental (symbols) and modelling (solid lines for best fit, and
dashed lines for 95 % confidence intervals) results for gas formation during
the simultaneous direct liquefaction and hydrodeoxygenation of beech saw- Figure 4. Influence of the plateau temperature on the product distribution
dust over the sulfide form of NiMo/Al2O3 at 350 8C, heating rate of after 60 min of simultaneous liquefaction and HDO of fine beech sawdust in
7.5 K min¢1, total pressure of 5 MPa, stirring speed of 500 rpm, sawdust-to- tetralin under 5 MPa hydrogen pressure over the sulfide form of NiMo/Al2O3,
tetralin mass ratio of 1:4 and mass of catalyst divided by the initial mass of total hydrogen pressure of 5 MPa, heating rate of 7.5 K min¢1, stirring speed
sawdust of 0.10. of 500 rpm, sawdust-to-tetralin mass ratio of 1:4 and mass of catalyst divid-
ed by the initial mass of sawdust of 0.10.

that solid sawdust is the main precursor for CO2, CO and CH4
formation. CO2 is the primary gaseous product that is formed, By elevating the reaction temperature from 300 to 350 8C,
which is advantageous in terms of deoxygenation and higher the rate of sawdust liquefaction apparently increased faster in
product quality and yield, as two oxygen atoms are removed comparison to char formation as the cumulative yield of solid
by each gas molecule. The mass yields of CO and CH4 are 8.2 residue decreased from 28 to 17 wt %, whereas the char con-
and 12.5 times lower, respectively, which is in relatively good tent in the solid residue increased from 12 to 70 wt %. Even
agreement with the non-catalysed runs reported by Beauchet though the increase of the reaction temperature resulted in
et al.[9] higher rates of biomass liquefaction, a lower fraction of tar
The lumping reaction mechanism of the LIQ-HT in tetralin is phase was obtained in the product because the HDO rates in-
shown in Scheme 1 and was proposed according to the inter- crease even faster if we apply a higher temperature, which
pretation of the experimental results. Sawdust de-polymeri- thus consumes more of the tar phase and results in significant-
sation results in gas (CO2, CO and CH4) and tar formation, ly higher yields of the oil phase and water. The optimised reac-
whereas the liquefied tar phase is further hydrodeoxygenated tion rate constants (Table 3) confirm that Ea2 and Ea3 are signifi-
to water and oil (liquid HDO products) by two possible hydro- cantly higher than Ea1 (i.e., the reaction temperature has
gen sources: tetralin and H2. The mechanism of the char for- a greater influence on HDO rate constants (k2 and k3) in com-
mation from the oil phase is explained in detail in the follow- parison to that of liquefaction (k1)). The activation energies of
ing sections. tar-phase HDO (Ea2 and Ea3) are in the same range as the activa-
tion energy reported for the catalytic HDO of the ¢OH group
in the liquefied biomass over NiMo/Al2O3 in our previous
work.[10]

Table 3. Activation energies and apparent reaction rate constants at


Scheme 1. Reaction mechanism of the simultaneous liquefaction and hydro- 350 8C for LIQ-HT.
treatment of lignocellulosic biomass.
Reaction Ean [J mol¢1] kn(350 8C) [min¢1]
Effect of the process temperature 1 (4.3 œ 0.1) Õ 10 4
(3.34 œ 0.03) Õ 10¢2
2 (7.0 œ 0.8) Õ 104 (1.83 œ 0.02) Õ 10¢2
The reaction temperature variation between 300 and 350 8C in- 3 (7.9 œ 0.6) Õ 104 (1.24 œ 0.01) Õ 10¢2[a]
fluenced the product distribution after 60 min for the fine 4 (1.5 œ 0.7) Õ 104 (3.1 œ 0.1) Õ 10¢2
beech liquefaction and hydro-treatment in tetralin at various 4C n.a.[b] > 103
4T n.a.[b] (3.9 œ 0.2) Õ 10¢2 Õ k4T[c]
(plateau) temperatures, a hydrogen pressure of 5 MPa and 5 (5 œ 2) Õ 104 (7.8 œ 0.2) Õ 10¢3
using the sulfide form of the NiMo/Al2O3 catalyst (Figure 4). In 6 (5 œ 1) Õ 104 (9.1 œ 0.9) Õ 10¢4
accordance with the Arrhenius law, the reaction rates of lique- 7 (7 œ 3) Õ 104 (7.0 œ 0.8) Õ 10¢4
faction, HDO and gas and char formation increased with ele- [a] Actual unit is min¢1 bar¢0.3 as m = 0.3 (the value of 0.297 was deter-
vating temperature; specifically, the solid- and tar-phase yields mined by regression analysis). [b] Within the error of estimation for all
decreased, whereas the fraction of gas-phase and HDO prod- model predictions but determined as approximately the same for reac-
tions 4T and 4C. [c] Actual unit is min¢1 bar¢1.
ucts increased noticeably at elevated temperatures.

ChemCatChem 2016, 8, 180 – 191 www.chemcatchem.org 183 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
The influence of the reaction temperature on the particle
size distribution of the solid residue is shown in Figure 5. A
comparison with the particle size distribution of the feedstock
(Figure 1) shows that the mean grain size is reduced during
LIQ-HT at 300 8C from 33 to 21 mm, whereas an increase in the
reaction temperature leads not only to a lower mass of the res-
idue but also to a smaller mean grain size of only 16 mm. The
particle size distribution also shows that the largest particles of
fine beech sawdust (up to 75 mm) were reduced as well as the
biggest bits that remained in solid residue were approximately
two times smaller.
The Van Krevelen diagram (Figure 6) shows that the insolu-
ble residue obtained after simultaneous liquefaction and
hydro-treatment contained a significantly lower hydrogen and Figure 6. Van Krevelen diagram for the feedstock and insoluble residue ob-
especially oxygen content in comparison to the original feed- tained after LIQ-HT at noted temperatures under 5 MPa pressure and with
the NiMo/Al2O3 catalyst.
stock as the residue is located in the area of the diagram that
is characteristic for bituminous coal and biomass-derived char
obtained after pyrolysis.[11] The process temperature does not also agree quite well with the spectrum of commercially avail-
have a remarkable influence on the O/C ratio of the solid resi- able alkali lignin (Figure S3 a), and hence, it may be concluded
due, whereas it affects the dehydrogenation slightly as the H/C that char is formed during the LIQ-HT from the chemically un-
ratio is decreased as the temperature is increased. Diffuse re- stable compounds that constitute the oil phase. These compo-
flectance infrared Fourier transform (DRIFT) spectra (Support- nents are most probably the aromatics from lignin, as the
ing Information, Figure S3) also confirm that the chemical com- latter were reported to be responsible for char formation even
position of the solid residue differs from the input feedstock at significantly lower temperatures, although in the presence
and matches well to DRIFT spectra of the bio-char obtained by of hydrogen donor solvents.[3a]
Azargohar et al. during pyrolysis at 400–475 8C.[11b] The spectra

Figure 5. Particle size distribution of solid residue after 60 min of LIQ-HT at a) 300 8C and 5 MPa, b) 325 8C and 5 MPa, c) 350 8C and 5 MPa and d) 350 8C and
8 MPa.

ChemCatChem 2016, 8, 180 – 191 www.chemcatchem.org 184 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
The mass balance for the carbon source distributed among (and in turn dissociated) H2 on catalyst surface or also by im-
the LIQ-HT products for all three reaction temperatures is proving hydrogen transfer (the effect included implicitly in k3).
shown in Table S1. Values are based on experimental data, Further exposure of the deoxygenated oil phase to harsh
whereas the carbon content for the tar phase and HDO prod- hydro-processing conditions causes the formation of free radi-
ucts was calculated by assuming the same elemental composi- cals (reaction 4) as a result of C¢C or C¢H bond cleavage on
tion for the solid and de-polymerised biomass fraction, that is, a catalyst surface. These free radicals seek a stabilisation route
tar phase, and the stoichiometric reaction progression of tar- among the competing possibilities and can be saturated by
based phase HDO as explained in the description of the kinetic hydrogen radicals (reaction 4T) released by the hydrogen
model. An increase of the temperature from 300 to 350 8C re- donor solvent assisted by a high hydrogen pressure, which
sulted in a higher conversion of the carbon-based components causes free radical stabilisation and prevents the recombina-
from the solid phase into HDO products, whereas the carbon tion that leads to the undesired char formation (reac-
content of the undesired products (i.e., the sum of the gas tion 4C).[2, 12]
phase and solid residue) decreased at a higher temperature, The experiment under N2 revealed the rate of the HDO dic-
which confirmed that the operation at temperatures up to tated by the hydrogen radicals from tetralin in the absence H2 ;
350 8C can be considered to be feasible because the process HDO was slow and only comparable to the reaction at 300 8C
was still not considerably compromised with the (undesired) under 5 MPa of H2, whereas the lack of H2 lowered the selectiv-
conversion of the carbon energy carriers into the carbon-con- ity towards free-radical stabilisation in comparison to re-combi-
taining side products in this range. nation, and hence, the apparent rate of char formation was
high despite a significantly lower presence of oil phase (the
precursor for char). However, the char yield was still nearly two
Effect of the process pressure and gas
times lower with respect to the non-catalytic reaction at 350 8C
A variation of the hydrogen pressure from 0 to 8 MPa had by using a hydrocarbon solvent without the hydrogen donor
a noteworthy influence on the course of the HDO and char for- activity reported by Schwaiger et al.[13] The reactions under
mation at 350 8C. Therefore, hydrogen pressure or its absence only 2 MPa H2 showed a considerably increased overall HDO
affected char, tar and HDO product yields and ratio, and practi- rate (in comparison to the reactions performed under N2),
cally no effect on gas formation (Figure 7). There was also no which resulted in a three-times lower yield of tar and a notably
noticeable difference in the particle size distributions of solid higher yield of HDO products. A subsequent increase of the H2
residues for the experiments performed at 5 or 8 MPa pressure has a direct effect on the HDO rate (reaction 4); spe-
(Figure 5). cifically, the reaction order with respect to H2 pressure was de-
termined to be 0.3 according to regression analysis. The influ-
ence of H2 pressure to prevent char (solid phase) formation is
seen clearly in Figure 7 and this aspect is described well by the
kinetic model. Operation in the absence of H2 led to a carbon-
rich solid phase as a result of char formation and nearly a two-
times lower conversion of the carbon source into the HDO
products in comparison to the reaction under the H2 atmos-
phere within the same time period, whereas the gas-phase
yield was not affected significantly (Table S1).
Experiments at various pressures of H2 or its absence
showed no significant effect on the gas formation rate or se-
lectivity. This indicates that decarboxylation (DCX) and decar-
bonylation (DCN) reactions can be considered mainly as the
side reactions of the thermal de-polymerisation (liquid-phase
pyrolysis) of solid biomass (clearly also independent of H2 pres-
Figure 7. Influence of process gas and hydrogen pressure on the product
sure) rather than the oxygen-containing tar phase deoxygena-
distribution after 60 min of simultaneous liquefaction and HDO of fine
beech sawdust in tetralin at 350 8C with NiMo/Al2O3. tion reactions. In the latter case, DCX and DCN reactions would
compete against catalytic HDO, and therefore, a variable H2
pressure or its absence would affect HDO/DCX and HDO/DCN
Molecular (gaseous) hydrogen and the hydrogen radicals selectivity significantly, and consequently, the CO and CO2
generated from tetralin are a possible hydrogen source for yields, which was not the case. Nevertheless, the DCX and DCN
HDO. Although it is well known that H2 can re-hydrogenate of the tar-based phase cannot be excluded completely, but
naphthalene into tetralin and, therefore, enhance the HDO ac- this mechanism was clearly not dominant for CO and CO2 for-
tivity of tetralin, those two mechanisms were considered inde- mation in this work. Our results are in accordance with the
pendent for the sake of dissimilarity, in which k2 governs the work of Wang et al. who reported a negligible effect (even
HDO reaction with tetralin in the absence of molecular hydro- with a negative trend) of the elevated H2 pressure on the gas
gen, whereas reaction 3 controls the influence of H2 pressure formation during the sawdust decomposition process in tetra-
to enhance the HDO by reaction between tar and dissolved lin without any heterogeneous catalyst.[7a]

ChemCatChem 2016, 8, 180 – 191 www.chemcatchem.org 185 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
Effect of the catalyst type tested in a mixture with the same mass of NiMo/Al2O3 to eval-
uate whether a synergetic effect of the catalytic cracking and
The effect of the catalyst type on the product distribution is HDO can cause a lower amount of insoluble residue, yet this
shown in Figure 8. Experiments were conducted in tetralin at was not the case. Pd/Al2O3 showed some HDO activity; never-
theless, it was significantly lower in comparison to that of
NiMo/Al2O3. Although HDO activity could have been included
in the kinetic model for each catalyst individually, additional
experiments would have been required to obtain reliable ki-
netic parameters (i.e., activation energy).

Effect of the biomass type and granulation


The influence of the wood type with respect to the product
yield and distribution was investigated by using oak, fir and
coarse beech sawdust, which all had a relatively similar particle
size distribution and specific surface area, and the effect of par-
ticle size was evaluated by comparing the fine and coarse
beech sawdust LIQ-HT in tetralin at 350 8C under 5 MPa H2 and
with the NiMo/Al2O3 catalyst.
Figure 8. Influence of catalyst type on the product distribution after 60 min
Clearly, the particle size distribution affected the rate of liq-
of simultaneous liquefaction and HDO of fine beech sawdust in tetralin at
350 8C, and under 5 MPa of H2 pressure. uefaction as the higher yields of solid and tar phase reveal that
the liquefaction of biomass is still taking place after 60 min if
coarse beech sawdust is used (in comparison to the fine beech
350 8C for 60 min under 5 MPa of H2 with a catalyst loading of counterpart; Figure 9). A nearly two-times lower specific sur-
10 wt % with regard to the mass of fine beech sawdust.
Among all reactions involved in LIQ-HT, the catalyst type af-
fects only the rates of HDO and cracking, whereas gas forma-
tion is completely independent (Figure S4). The gas yield and
CO/CO2 ratio also corresponded well to the non-catalysed bio-
mass liquefaction in tetralin reported by Wang et al.,[7a] which
confirms that gas formation can be attributed predominantly
to the thermal de-polymerisation of biomass and not to the re-
actions promoted by heterogeneous catalysts. Therefore, the
presence of these catalysts neither promotes nor inhibits CO2
or CO formation as the probability of the precursor (biomass
particles) to be in physical contact with the catalyst surface is
insignificant. However, the catalyst type had some influence on
CH4 yield (Figure S4), which was the highest (but still as low as
1.5 wt %) if the HDO catalyst NiMo/Al2O3 was used and the
Figure 9. Influence of wood type and its particle size on the product distri-
lowest with zeolite Y with no HDO active sites. The difference bution after 60 min of simultaneous liquefaction and HDO in tetralin at
in CH4 yield might be caused by the HDO of components such 350 8C under 5 MPa of H2 pressure with NiMo/Al2O3.
as methanol, formaldehyde, formic acid, methyl ethers or
esters or methane formation from CO and CO2 as suggested
by Beauchet et al. who measured a similar selectivity for CO2, face area of coarse beech sawdust in comparison to the finer
CO and CH4 in the non-catalysed process at 330 8C but pointed fraction affects the rate of the solid–liquid interface reaction of
out that the presence of Raney Ni had a significant effect on liquefaction directly. A lower biomass conversion is also reflect-
the selectivity of gas formation (CH4 selectivity increased from ed in a decreased yield of HDO products, although the experi-
a few to nearly 50 %) but not on the global yield as a result of mental content values for tar and HDO products might not be
methane formation from the evolving CO and CO2 gases and very precise as a result of fractionating (method), whereas the
not hydrogenolysis.[9] The catalysts tested in this work were absolute experimental errors for the solid- and gas-phase
clearly much less active for the methane formation reactions. yields are expected to be much lower (below 2 wt %).
As the catalysts primarily only promoted HDO reactions, it is A very similar conversion of sawdust into HDO products was
of no surprise that the highest yield of HDO products was ob- obtained with the use of fine beech, coarse oak and fir saw-
tained by NiMo/Al2O3, a catalyst developed for hydro-treat- dust (Figure 9). In contrast to catalyst or H2 pressure variation,
ment. Zeolite Y showed an insignificant HDO activity, and con- a difference in wood type had little, but still noticeable, effect
sequently, the yield of the polar tar phase was the highest and on the gas-phase composition formed during the LIQ-HT in
that of the HDO products was the lowest. Zeolite Y was also tetralin, which proves once again that gas was formed directly

ChemCatChem 2016, 8, 180 – 191 www.chemcatchem.org 186 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
acidified polyol in comparison to phenol and especially tetralin
is in accordance with previous reports.[3b]
The progression of solvolysis and HDO in glycerol can be fol-
lowed by the FTIR spectroscopy of liquid samples as the prod-
ucts of biomass de-polymerisation were soluble in the liquid
phase. The FTIR spectra shown in Figure 11 correspond to the

Figure 11. FTIR spectra of the liquid phase during LIQ-HDO in glycerol and
Figure 10. Influence of sawdust type and reaction temperature on the gas-
glycerol, acidified by p-TSA, at 350 8C and 5 MPa hydrogen pressure over
phase composition in the reactor at a particular time (other characteristics
NiMo/Al2O3.
noted in the caption of Figure 9).

mechanism of the biomass solvolysis by polyhydroxyl alcohols


from biomass during the non-catalytic thermal de-polymeri- reported previously,[14] in which the glycoside bond of cellulose
sation (Figure 10). The differences in the gas-phase composi- or hemicellulose is de-polymerised followed by furanose or
tion were probably low because of the similar cellulose, hemi- pyranose ring opening and the formation of levulinates with
cellulose and lignin content in the tested wood samples. In the the characteristic peaks located at ñ = 1715–1720 cm¢1, where-
LIQ-HT of the pure lignin in tetralin at 400 8C under H2 pressure as carbonyl groups conjugated to aromatic rings showed char-
and in the presence of CoMo/Al2O3, Araya et al. reported acteristic signals at ñ = 1640 cm¢1. Evidence of deoxygenation
a 2.5 times lower gas yield in comparison to pure cellulose, is also noticeable in the spectra shown in Figure 11 by the di-
whereas the selectivity towards gaseous hydrocarbons was minishing areas of the peaks at ñ = 3300 and 1040–1117 cm¢1
50.3 wt % for lignin and only 15.3 wt % for cellulose.[8] that represent O¢H and C¢O groups, respectively. The FTIR
spectra of liquid-phase samples collected during LIQ-HT with
the use of phenol or especially tetralin do not reveal much in-
Effect of the solvent
formation about the reactions for two reasons: i) an intense
The influence of the solvent type was compared directly in signal of aromatics masks the FTIR spectra and ii) the insolubili-
terms of solid residue as the solubility of de-polymerised or hy- ty of the polar tar phase in non-polar tetralin causes most of
drodeoxygenated components in solvents is significantly differ- the polar phase to be adsorbed physically or chemically on the
ent for non-polar tetralin, semi-polar phenol and polar glycer- solid residue, polar alumina or zeolite, and only the completely
ol; in addition, CO and CO2 can also be formed from an deoxygenated components have a low enough polarity to be
oxygen-containing solvent. Liquefaction in glycerol is slower in extracted by a non-polar solvent; however, the C¢C and C¢H
comparison to that in tetralin, whereas the addition of a homo- signals of those components are clearly not dominant enough
geneous acidic catalyst (p-toluenesulfonic acid; p-TSA) im- to be seen in the spectra or are masked by the C¢C and C¢H
proves liquefaction significantly as the remaining solid residue signal of the solvent (Figure S6).
was only 1 wt % of the initial sawdust mass (Figure S5). The
acidic nature of phenol does not require an additional homo-
Conclusions
geneous catalyst for a relatively high biomass conversion, and
the amount of gases formed is lower in comparison to that in Simultaneous direct liquefaction and hydrodeoxygenation of
the reaction with tetralin, which is mostly in the form of CO2. lignocellulosic biomass in hydrogen donor solvents involves
Therefore, the formed gases result from biomass decomposi- several reaction steps, which include biomass de-polymeri-
tion and not from the phenolic hydroxyl group. In contrast, sation, decarboxylation or decarbonylation, char formation and
the use of glycerol results in extensive gas formation, which is hydrodeoxygenation (HDO), governed by hydrogen supplied
more than that generated from the decarboxylation and decar- in gaseous form or generated in situ. A reaction mechanism
bonylation of the solvent. A higher liquefaction activity of the was proposed, and a new lumped kinetic model was devel-

ChemCatChem 2016, 8, 180 – 191 www.chemcatchem.org 187 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
oped according to the results obtained by varying the reaction The total pore volume was determined by estimating the amount
time, temperature, pressure, process gas, catalyst type, wood of N2 adsorbed at the relative pressure (P/P0) of 0.95, and the Hor-
type and solvent. Temperature affected all of the investigated vath–Kawazoe method was used to determine the pore size and
pore volume determination for micro-porous zeolite Y. Information
reactions according to the Arrhenius law, and a temperature
about the catalyst compositions is given according to the data pro-
increase from 300 to 350 8C resulted in an approximately
vided by the suppliers.
2.5 times higher yield of HDO products, whereas the mass of
solid residue decreased by 39 %. The significantly higher activa-
tion energy of the HDO reactions in comparison to biomass
Materials
liquefaction caused the lower yields of the tar phase at higher
temperatures, for which the latter is the product of liquefac- Sawdust samples from debarked European beech (Fagus sylvatica),
tion and the precursor for HDO, as well as higher correspond- European silver fir (Abies alba) and sessile oak (Quercus petrea) cut
down in the forests near Postojna, Slovenia (458 41’ N, 148 12’ E)
ing oil yields. However, an increase of the temperature to over
were dried on a metal tray (the layer thickness below 3 mm) at
350 8C might result in lower oil yields because of char forma-
80 8C under a N2 flow for 24 h and kept in a desiccator at RT
tion, especially at reaction times longer than 60 min. This study before use. Two different particle size fractions of European beech
also showed that hydrogen pressure affected the HDO rate were tested. The dried wood samples were characterised by field
positively (with an apparent reaction order of 0.3) and it also emission scanning electron microscopy (FE-SEM), BET, DRIFTS and
inhibited char formation considerably, presumably by saturat- Pregl–Dumas (CHNS-O) methods.
ing the free radicals formed during catalytic cracking, and thus All solvents and gases were obtained from commercial suppliers.
prevented their polymerisation. As the formation of the polar Tetralin (Š 97 wt %), phenol (Š 99 wt %) and p-toluenesulfonic acid
tar and gas phase is a process caused by the thermal degrada- (Š 98 wt %) were from Sigma–Aldrich (St. Louis, MO, USA). Glycerol
tion and de-polymerisation of lignocellulosic biomass, the cata- (Š 99 wt %) was from Pharmachem Sušnik (Ljubljana, Slovenia). H2
lyst type did not affect the rate of tar formation and the com- (5.0) and N2 (5.0) were from Messer (Bad Soden am Taunus, Germa-
position of the latter, whereas it has a considerable effect on ny).
HDO, and hence the highest oil and the lowest tar yields were
obtained by using the sulfide form of NiMo/Al2O3 followed by
Pd/Al2O3, which also showed some HDO activity. The insignifi- Simultaneous liquefaction and hydro-treatment
cant HDO activity of zeolite Y resulted in the highest solid- The simultaneous liquefaction and catalytic hydro-treatment of
phase conversion and tar-phase yield, whereas the fraction of sawdust was investigated by using a 300 mL cylindrical stainless-
HDO products was the lowest. As already mentioned in the lit- steel slurry reactor (Parker Autoclave Engineers, Erie, PA, USA). The
erature and confirmed in this study, there are other solvents same reactor configuration was used as reported previously,[4b, c] in
available for the even more efficient biomass liquefaction and which further details about the experimental setup can be found
further hydro-processing in comparison to rather expensive (Figure S7).
tetralin; nonetheless, most of them contain the undesired The reaction mixture contained a dried sawdust sample (20 g) and
chemically bonded oxygen atom that reduces the gross calorif- a solvent (tetralin, phenol or glycerol) in a 1:4 mass ratio, and the
ic value of the resulting product and its applicability and po- catalyst mass (after activation) was set to 10 wt % with regard to
tential use as a transportation fuel. the initial mass of dry wood. After the reactor vessel was attached
to its casing, the headspace was pressurised and depressurised
with H2 (or N2 if it was used as the process gas) twice to displace
Experimental Section air before the final pressure was set to 2, 5 or 8 MPa at a constant
gas flow through the reactor (1000 mLN min¢1). The reactor was op-
Catalyst activation and characterisation erated in a batch regime if we consider the solid and liquid phase,
and a continuous H2 flow was established to ensure sufficient and
Three catalyst types that are used commonly on an industrial scale
constant H2 availability throughout the process and to enable
were tested in this study (Table 1). Commercially available NiMo/
online gas-phase analysis. The agitation speed was set to
Al2O3 (quadri-lobe-shaped) catalyst was received in oxide form and
500 min¢1, and the reactor content was heated (7.5 K min¢1) to the
was sent to Eurecat France (La Voulte-sur-Rhúne, France) for the
desired temperature (300, 325 or 350 8C), which was kept constant
sulfide form rendition by the Totsucat process using H2S/H2 sulfide
for 60 min except if noted otherwise, followed by H2 supply shut-
agent. Palladium supported on 2–4 mm Al2O3 spheres (G-68 C,
down and rapid reactor cooling below 30 8C. After depressurisa-
Sìd-Chemie, Munich, Germany), was dried under N2 at 250 8C for
tion, the reactor was flushed with N2, opened and weighed.
60 min and reduced subsequently under 500 mL min¢1 H2 flow in
a tube furnace before use. The initial reduction temperature was
set to 150 8C and kept constant for 1 h followed by gradual heat-
ing (0.25 K min¢1) to 350 8C, at which the Pd-containing catalyst Gas-, liquid- and solid-phase analysis
was reduced for an additional 7 h. Zeolite Y (CBV 300, Zeolyst, Con-
The cooled and decompressed gas from the reactor was analysed
shohocken, PA, USA) was used in powder form and was activated
on-line by FTIR spectroscopy (Spectrum 100, PerkinElmer, Waltham,
by calcination in air by using a ventilated furnace at 500 8C for 5 h
MA, USA) in the range of 450–4000 cm¢1, and the detected compo-
to substitute ammonium cations with H+.
nents were quantified. Six liquid samples of 0.5 mL were with-
The specific surface area of the catalysts (as well as sawdust sam- drawn and stored during the LIQ-HT experiments, subsequently ap-
ples) was determined by the BET N2 physisorption method by plied to a silicon wafer and analysed off-line. Liquid- and gas-phase
using an ASAP 2020 (Micrometrics, Norcross, GA, USA) instrument. analyses, as well as the component or functional group quantifica-

ChemCatChem 2016, 8, 180 – 191 www.chemcatchem.org 188 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
tion methodology by FTIR is described in detail in our previous wood particles spread over dark (conductive) graphite tape. The
studies.[4a, 10] particle sizes were subsequently determined by using image pro-
cessing and analysis software (ImageJ, Bethesda, MA, USA), in
After each run with tetralin, the LIQ-HT products were determined
which 400–1000 particles were digitally characterised in the SEM
and quantified according to the sequence shown in Scheme 2. The
images for each sample. The minimal Feret diameter was consid-
liquid phase was separated from the solid and tar counterparts by
ered as the most relevant particle characteristic to simulate sieve
analysis, and we assumed that the observed particles are isotropic
and similar to the solid of revolution obtained by rotating the par-
ticles’ 2 D projection around the Feret axis. The solid of revolution,
the 2 D projection of which has a minimal Feret diameter lower
than the mesh opening, can theoretically pass through the respec-
tive sieve.
The FTIR spectra of sawdust and insoluble residue were recorded
using the DRIFTS method by averaging 16 scans with the resolu-
tion of 4 cm¢1 by using a DiffusIR (Pike Technologies, Madison, WI,
Scheme 2. Separation and analysis of LIQ-HT products.
USA) environmental chamber flushed with nitrogen installed in the
spectrometer. The KBr background was recorded before the solid
filtration, and the reactor vessel and filter cake were rinsed with samples were diluted with KBr in a 1:2 mass ratio, placed in
toluene several times; subsequently, the residue (insoluble in non- a sample cup (10 mm diameter, 2.3 mm deep) and analysed. The
polar solvent) was dried at 110 8C overnight (16 h) and weighed, elemental (CHNS-O) composition of insoluble residue was deter-
and the filtrate, which contained solvent and the oil phase, was mined by the Pregl–Dumas method by using a vario EL cube (Ele-
centrifuged for 10 min at 7500 rpm and examined visually for the mentar, Hanau, Germany) analyser.
potential presence of a polar phase. Polar tarry components (T)
were extracted from the filter cake with absolute ethanol. The in-
soluble residue (S) and solid catalyst were separated from the dilut- Kinetic model of simultaneous liquefaction and hydro-treat-
ed tar phase by two-step filtration (0.8 mm mesh to separate cata- ment in tetralin
lyst particles and a black ribbon filter paper for the remaining solid
The lumped kinetic model was based on the reaction mechanism
phase), rinsed thoroughly with absolute ethanol, and dried over-
(Scheme 1) developed according to the experimental results and
night (16 h) at 80 8C. If the catalyst could not be separated from
modelling responses for various reaction pathways.
the insoluble residue (i.e., if zeolite powder was used), the mass of
the latter was calculated by the subtraction of the initial catalyst The reaction mechanism includes the reactions of liquefaction and
mass from the mass of the dried filter cake after extraction liquid-phase pyrolysis to partly de-polymerised lignocellulosic
[Eqs. (1)–(5)]. macro-molecules (k1), accompanied by gas formation (k6–k8), which
became extensive if temperatures exceeded 280 8C [Eqs. (6)–(10)].
wS ¼ mS =m0 ð1Þ
dww ðtÞ
wT ¼ mT =m0 ð2Þ ¼ ¢k1 ¡ ww ðtÞ¢k 5 ¡ ww ðtÞ¢k 6 ¡ ww ðtÞ¢k7 ¡ w w ðtÞ ð6Þ
dt
X
L dwCO2 ðtÞ
wG ¼ mG =m0 ¼ ðni ¡ Mi Þ=m0 ð3Þ ¼ k 5 ¡ ww ðtÞ ð7Þ
dt
l¼1
dwCO ðtÞ
wHDO ¼ 1¢wS ¢wT ¢w G ð4Þ ¼ k 6 ¡ w w ðtÞ ð8Þ
dt
wOIL ¼ wHDO ¢wH2 O ¼ wHDO ¢mH2 O =m0 ð5Þ dwCH4 ðtÞ
¼ k 7 ¡ ww ðtÞ ð9Þ
dt
The yield of solid residue (wS) or tar phase (wT) was calculated by w G ðtÞ ¼ wCO2 ðtÞ þ wCO ðtÞ þ w CH4 ðtÞ ð10Þ
dividing the mass of the solid residue (mS) or tar phase (mT) by the
initial mass of sawdust (m0). The yield and mass of the formed per-
The hydrodeoxygenation of the liquefied (tarlike) biomass phase
manent gases (wG and mG) was calculated according to the on-line
proceeded through two different mechanisms: i) transfer hydro-
analysis of the main components (l), in which L is their number, de-
(deoxy)genation caused by the hydrogen radicals formed by tetra-
tected in the gas phase during reactions and depressurisation. If
lin (extent of this effect being determined by the reaction under N2
we consider that the tar phase is the main product of liquefaction
pressure) and ii) hydrogenolysis caused by H2. The experiments at
and that it is subsequently hydrodeoxygenated, then the HDO
different hydrogen pressures (PH2 ) showed that its impact on reac-
product yield (wHDO) can be calculated as the balance [Eq. (4)], and
tion rate was lower than that encompassed by the reaction order
this fraction contains the completely deoxygenated oil phase (mOIL)
of unity; therefore, the reaction order (m) with respect to pressure
dissolved in a non-polar solvent (only C¢H and C¢C bonds were
was determined by regression analysis [Eq. (11)].
present in the liquid phase after the HDO with tetralin as the sol-
vent) and water (mH2 O ) [Eq. (5)] that is separated from the gas dwT ðtÞ
phase and collected in a condenser. ¼ k1 ¡ ww ðtÞ¢k2 ¡ wT ðtÞ¢k3 ¡ wT ðtÞ ¡ Pm ð11Þ
dt H2

The particle size distribution for all sawdust samples (and the solid
residue after reactions) was determined by FE-SEM (SUPRA 35 VP, During HDO, water and the deoxygenated oil phase were consid-
Carl Zeiss, Jena, Germany) operated at a high acceleration voltage ered to be formed in the mass ratio of 2:3 according to the gener-
(20 kV) to ensure a sufficient contrast of bright (non-conductive) alised stoichiometry. If we assumed that the de-polymerised tar

ChemCatChem 2016, 8, 180 – 191 www.chemcatchem.org 189 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
phase had the same elemental composition as sawdust, the stoi- was modified to minimise the mutual correlation between the acti-
chiometric ratio of 0.417:0.583 was determined for the hydrodeox- vation energy and pre-exponential factor during regression analy-
ygenation of the treated feedstock. Water-to-oil mass ratios be- sis.
tween 0.40:0.60 and 0.45:0.55 were tested, and it was determined
Regression analysis and modelling in general was performed by
that the latter also allowed for the carbon and oxygen mass bal-
using Matlab 7.12.0. For optimisation purposes, the Nelder–Mead
ance closure for the LIQ-HDO process. As a result of the experi-
method was applied for the initial optimisation of the apparent re-
mental uncertainty and non-ideal stoichiometry of actual deoxyge-
action rate constants followed by the Levenberg–Marquardt
nation reactions, this ratio can only be given as accurately as
method for the final Ean and kn(TP) optimisation [the minimisation
0.4:0.6, whereas the balances for water and oil may thus be formu-
of the objective function given by Eq. (20)] and the Jacobian
lated as according to Equations (12) and (13).
matrix computation required for the subsequent determination of
dwH2 O ðtÞ the 95 % confidence intervals for non-linear least squares kinetic
¼ 0:4 ¡ ðk2 ¡ w T ðtÞ þ k3 ¡ wT ðtÞ ¡ Pm Þ ð12Þ parameter estimation.
dt H 2

dwOIL ðtÞ f ðE a1 ,k1 ðT p Þ, . . . ,E a7 ,k 7 ðT p Þ,k4C =k4T , mÞ ¼


¼ 0:6 ¡ ðk2 ¡ wT ðtÞ þ k 3 ¡ wT ðtÞ ¡ Pm Þ¢k4 ¡ wOIL ðtÞþ
dt H 2
ð13Þ
X
J XI
ðk4T ¡ PH2 Þ ¡ w OIL* ðwi,j meas ¢w i,j calc ðE a1 ,k1 ðT P Þ, . . . ,E a7 ,k 7 ðT P Þ,k4C =k4T , mÞÞ2
j¼1 i¼1

Char formation is triggered by the free radical (OIL*) formation (re- ð20Þ
action 4) after the compounds that constitute oil are cracked on
acidic catalyst support. Free radicals can either re-combine to form The measured mass yields (wi,jmeas) of each component i (I is their
high-molecular-weight compounds (char; reaction 4C) or can be number) were fitted to modelled values (wi,jcalc) simultaneously for
stabilised by a radical scavenger that saturates free radicals and all LIQ-HT runs (j; J is their number) that were included in model-
prevents the re-combination [reaction 4T; Eqs. (14) and (15)]. ling study to obtain the apparent kinetic parameters for liquefac-
tion, HDO, char and gas formation during the LIQ-HT of beech
dwOIL* ðtÞ
¼ k4 ¡ wOIL ðtÞ¢k 4C ¡ w OIL* ðtÞ¢ðk 4T ¡ PH2 Þ ¡ w OIL* ðtÞ ð14Þ wood in tetralin over NiMo/Al2O3.
dt
dwCHAR ðtÞ
¼ k4C ¡ w OIL* ðtÞ ð15Þ
dt
Acknowledgements
The rates of the reactions between free radicals (reaction 4T and
4C) are orders of magnitude higher than the rate of free radicals The authors gratefully acknowledge the financial support of the
formation (reaction 4), therefore, the following applies: i) The con- Slovenian Research Agency (ARRS) through the Programme P2-
centration of free radicals is negligible. ii) The constant k4 is the 0152. The authors also acknowledge Gozdno Gospodarstvo Post-
rate-determining step parameter for charring, and the k4C/(k4T·PH2 + ojna for wood samples and Eva Kovič for her excellent laboratory
k4C) and k4T·PH2 /(k4T·PH2 + k4C) ratios determine the selectivity to- work. Special thanks go to Eurecat for the catalyst sulfide from
wards either char formation or radical stabilisation. In this work, rendition by the Totsucat process.
the selectivity was considered to be temperature independent as
indicated by the measured data. As the experimental results
showed a positive influence of higher hydrogen pressure to pre- Keywords: biomass · kinetics · reaction mechanisms ·
vent char formation, the effect of hydrogen pressure (PH2 ) was also solvolysis · supported catalysts
considered in the respective backward reaction.
This simplified the balance for oil and char and eliminated the bal-
[1] a) M. Grilc, B. Likozar, J. Levec, Catal. Today 2015, 256, 302 – 314; b) N.
ance for free radicals [OIL*; Eqs. (16)–(18)]. Schwaiger, R. Feiner, H. Pucher, L. Ellmaier, J. Ritzberger, K. Treusch, P.
Pucher, M. Siebenhofer, Chem. Ing. Tech. 2015, 87, 803 – 809; c) C. De-
dwOIL ðtÞ scorme, P. Gallezot, C. Geantet, C. George, ChemCatChem 2012, 4,
¼ 0:6 ¡ ðk2 ¡ wT ðtÞ þ k 3 ¡ wT ðtÞ ¡ Pm Þ¢
dt H2
ð16Þ 1897 – 1906; d) H. Pucher, N. Schwaiger, R. Feiner, L. Ellmaier, P. Pucher,
k4 ¡ ðk 4C =ðk4T ¡ PH2 þ k4C ÞÞ ¡ wOIL ðtÞ B. S. Chernev, M. Siebenhofer, Green Chem. 2015, 17, 1291 – 1298.
[2] a) N. P. Vasilakos, D. M. Austgen, Ind. Eng. Chem. Process Des. Dev. 1985,
dwCHAR ðtÞ 24, 304 – 311; b) E. Churin, P. Grange, B. Delmon, R. Commission of the
¼ k4 ¡ ðk4C =ðk4T ¡ PH2 þ k4C ÞÞ ¡ wOIL ðtÞ ð17Þ
dt European Communities. Directorate-General for Science Development,
Quality improvement of pyrolysis oils: final report, Commission of the Eu-
wS ðtÞ ¼ w W ðtÞþwCHAR ðtÞ ð18Þ ropean Communities, 1989.
[3] a) A. Toledano, L. Serrano, J. Labidi, A. Pineda, A. M. Balu, R. Luque,
Apparent reaction rate constants were for the corresponding tem- ChemCatChem 2013, 5, 977 – 985; b) S.-A. Rezzoug, R. Capart, Biomass
perature at each time increment calculated according to the modi- Bioenergy 1996, 11, 343 – 352.
[4] a) M. Grilc, B. Likozar, J. Levec, Appl. Catal. B 2014, 150 – 151, 275 – 287;
fied Arrhenius equation [Eq. (19)]:
b) M. Grilc, G. Veryasov, B. Likozar, A. Jesih, J. Levec, Appl. Catal. B 2015,
  ŒŒ 163, 467 – 477; c) T. B. Čelič, M. Grilc, B. Likozar, N. N. Tušar, ChemSus-
E 1 1
k n ðTðtÞÞ ¼ k n ðT P Þ ¡ exp ¢ a n ¢ ð19Þ Chem 2015, 8, 1703 – 1710; d) A. R. Ardiyanti, A. Gutierrez, M. L. Honkela,
R TðtÞ T P A. O. I. Krause, H. J. Heeres, Appl. Catal. A 2011, 407, 56 – 66; e) E. Furim-
sky, Appl. Catal. A 2000, 199, 147 – 190.
in which kn(TP) is the apparent rate constant for reaction n at the [5] S.-A. Rezzoug, R. Capart, Appl. Energy 2002, 72, 631 – 644.
temperature of 623 K (TP), which was chosen as the temperature of [6] J. D. Rocha, C. A. Luengo, C. E. Snape, Renewable Energy 1996, 9, 950 –
plateau set for the most of experiments. The Arrhenius equation 953.

ChemCatChem 2016, 8, 180 – 191 www.chemcatchem.org 190 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
[7] a) G. Wang, W. Li, H. Chen, B. Li, Energy Sources Part A 2007, 29, 1221 – [12] D. Liu, L. Song, P. Wu, Y. Liu, Q. Li, Z. Yan, Bioresour. Technol. 2014, 155,
1231; b) Y. Yan, J. Xu, T. Li, Z. Ren, Fuel Process. Technol. 1999, 60, 135 – 152 – 160.
143. [13] N. Schwaiger, R. Feiner, K. Zahel, A. Pieber, V. Witek, P. Pucher, E. Ahn, P.
[8] P. E. Araya, S. E. Droguett, H. J. Neuburg, R. Badilla-Ohlbaum, Can. J. Wilhelm, B. Chernev, H. Schrçttner, M. Siebenhofer, Bioenergy Res. 2011,
Chem. Eng. 1986, 64, 775 – 780. 4, 294 – 302.
[9] R. Beauchet, L. Pinard, D. Kpogbemabou, J. Laduranty, L. Lemee, J. L. [14] a) T. Yamada, H. Ono, J. Wood Sci. 2001, 47, 458 – 464; b) B. Girisuta,
Lemberton, F. Bataille, P. Magnoux, A. Ambles, J. Barbier, Bioresour. Tech- L. P. B. M. Janssen, H. J. Heeres, Ind. Eng. Chem. Res. 2007, 46, 1696 –
nol. 2011, 102, 6200 – 6207. 1708.
[10] M. Grilc, B. Likozar, J. Levec, Biomass Bioenergy 2014, 63, 300 – 312.
[11] a) H. S. Kambo, A. Dutta, Renewable Sustainable Energy Rev. 2015, 45,
359 – 378; b) R. Azargohar, S. Nanda, J. A. Kozinski, A. K. Dalai, R. Sutarto, Received: July 28, 2015
Fuel 2014, 125, 90 – 100. Published online on December 2, 2015

ChemCatChem 2016, 8, 180 – 191 www.chemcatchem.org 191 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like