Anda di halaman 1dari 6

Bioresource Technology 127 (2013) 306–311

Contents lists available at SciVerse ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Hydrotreatment of bio-oil over Ni-based catalyst


Xinghua Zhang a,b, Tiejun Wang a,⇑, Longlong Ma a,⇑, Qi Zhang a, Ting Jiang a,b
a
Key Laboratory of Renewable Energy and Gas Hydrate, Guangzhou Institute of Energy Conversion, Chinese Academy of Sciences, Guangzhou 510640, PR China
b
Graduate School of Chinese Academy of Science, Beijing 100049, PR China

h i g h l i g h t s

" Ni/HZSM-5 catalyst possesses the catalytic activity for cracking and hydrogenation.
" Cracking reaction and hydrogenation reaction take place simultaneity during the hydrotreating process.
" Reaction temperature has great effect on product selectivity for benzene and cyclohexane.
" The pH value of bio-oil increased from 2.27 to 4.07 after upgrading.

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

Article history: Inexpensive non-sulfided Ni-based catalysts were evaluated for hydrotreatments using phenol as
Received 5 January 2012 model compound. HZSM-5, a zeolite with different ratio of Si/Al and c-Al2O3 were impregnated with
Received in revised form 2 July 2012 Ni(NO3)26H2O and calcined at 450 °C. Conversion rates and product distribution for treatment of phenol
Accepted 30 July 2012
at 160–240 °C in the presence of catalysts with nickel loads of 6, 10, 14 and 17 wt.% were determined.
Available online 30 August 2012
Phenol conversion was highest (91.8%) at 240 °C in the presence of HZSM-5(Si/Al = 38) loaded with
10% Ni. When hydrotreatment was carried out with bio-oil obtained from pyrolysis of pine sawdust
Keywords:
under the optimal conditions determined for phenol, the pH of bio-oil increased from 2.27 to 4.07, and
Bio-oil
Phenol
the hydrogen content increased from 6.28 to 7.01 wt.%. The decrease in acidity is desirable for the use
Ni-based catalysts of upgraded bio-oil.
Hydrodeoxygenation (HDO) Ó 2012 Published by Elsevier Ltd.
Upgrading

1. Introduction Gandarias et al., 2008; Senol et al., 2007). However, such HDS
catalysts are not desirable for bio-oil hydrotreating because HDS
Biomass is a promising alternative energy source owing to its catalysts require the addition of sulfur-containing compounds,
renewability and zero emission (Zhang et al., 2005). Fast pyrolysis such as H2S or thiophene, in the reaction zone to keep the catalysts
is the leading method for conversion of biomass to liquid products, active, which increase the contamination risk of bio-oil (Bui et al.,
known as bio-oils; however, these oils cannot be directly used as 2011) and because water in bio-oil can react with alumina (Laurent
transportation fuels due to their high oxygen and water content and Delmon, 1994; Zakzeski et al., 2010). For example, alumina is
(Fisk et al., 2009; Peng et al., 2009). The oxygenated compounds known to be metastable under hydrothermal conditions and can
in bio-oil cause high viscosity, poor thermal and chemical stability, partially transform into boehmite under processing conditions. In
corrosion and immiscibility with hydrocarbon fuels (Gong et al., addition, pyrolysis oil has been upgraded by high pressure thermal
2011; Scholze et al., 2001; Vitolo et al., 2001;Yang et al., 2009).Con- treatment (Mercader et al., 2010a), followed by co-processing in
sequently, upgrading is needed to reduce the oxygen content. standard refinery units (Mercader et al., 2010b). Catalysts based
Hydrodeoxygenation (HDO) is considered the most effective on Pd and Pt have also been studied due to their high selectivity
method for bio-oil upgrading (Ardiyanti et al., 2011; Gandarias towards alkanes (Bejblová et al., 2005; Díaz et al., 2007; Fisk
et al., 2008; Yang et al., 2009; Zhao et al., 2009) and hydrodesulfu- et al., 2009; Zhao et al., 2009; Oasmaa et al., 2010), but the treat-
rization (HDS) catalysts such as sulfided Co–Mo and Ni–Mo sup- ments were carried out at over 350 °C, facilitating production of
ported on c-Al2O3 have been employed (Bunch et al., 2007; polymer and coke-like products and causing catalyst deactivation
(Adjaye and Bakhshi, 1995; Vitolo et al., 2001; Yang et al., 2009).
Furthermore, since HDO of bioliquids is expected to be a large-
⇑ Corresponding authors. Tel.: +86 20 87057673; fax: +86 20 87057737. scale process, employment of noble metal-based catalysts could
E-mail addresses: wangtj@ms.giec.ac.cn (T. Wang), mall@ms.giec.ac.cn (L. Ma).

0960-8524/$ - see front matter Ó 2012 Published by Elsevier Ltd.


http://dx.doi.org/10.1016/j.biortech.2012.07.119
X. Zhang et al. / Bioresource Technology 127 (2013) 306–311 307

significantly raise production costs. Thus, it seems more reasonable 2.4. Products analysis
to use Ni-based catalysts for bio-oil HDO due to their high activity
for hydrogenation and lower cost. Products collected in vacuum plastic bags were analyzed by gas
In order to reduce the oxygen content of bio oil at temperature chromatography (GC) (Shimadzu GC2010 with a flame ionization
of 240 °C, inexpensive non-sulfided Ni-based bifunctional catalysts detector (FID) and a DB-5 column, 30 m  0.25 mm  0.1 lm, N2
were prepared. The activity of the catalysts was evaluated in water as carrier gas with 99.995% purity), with commercial standard re-
using phenol as a model compound. The reaction parameters were agents as external standards. The vaporization temperature was
optimized with a focus on the effects of supporting material, reac- 100 °C, the detector temperature was 200 °C, and the oven temper-
tion temperature, and nickel loading. Furthermore, upgrading of ature program was increased from 50 °C to 120 °C at a rate of 5 °C/
bio-oil through hydrotreating was carried out under the optimized min.
conditions. Conversions of phenol and product selectivity were calculated
based on the formulas:
P
2. Methods moles C p
Conversion % ¼  100
molesphenolfed
2.1. Preparation of catalysts
moles C p
Ni/HZSM-5 catalysts were prepared through wet impregnating C p selectivity % ¼ P  100
moles C p
of different amounts of HZSM-5 (prepared by Na-ZSM-5 zeolite
with 1 mol/L NH4NO3 solution by ion exchange at 90 °C) with Ni Cp represents the contents of benzene, cyclohexane, cyclohex-
(NO3)26H2O (A.R. grade) aqueous solutions. The solutions were ene, cyclohexanol and methyl-cyclopentane, respectively.
evaporated and the residues were dried at 120 °C, then calcined The water content of the bio-oil was determined by Karl Fischer
at 450 °C in air. Catalysts containing 6, 10, 14 and 17 wt.% Ni were titration (ASTM D 1744, GB11146-89) with a Metrohm 787 KF
prepared. In the same way, the 10 wt.% Ni/c-Al2O3 catalysts were Titrino. The acidity was evaluated with a PHC-3C precision pH-me-
prepared. c-Al2O3 was from the Nankai University Catalyst ter from Shanghai REX Instrument Factory. The elemental analysis
Co. Ltd., (Tianjin, China). was carried out with a VARIO ELI Elemental Analyzer (Elementar
Analysensysteme Gmbh, Germany). The composition of the bio-
oil was determined by a Shimadzu GC–MS-QP2010 equipped with
2.2. Hydrotreatments a column of DB-5 (30 m  0.25 mm  0.1 lm). The carrier gas was
He with 99.995% purity, and the oven temperature program ranged
The aqueous-phase catalytic reactions of phenol were carried from 40 °C (holding for 5 min) to 280 °C (holding for 10 min) at a
out in a 100-ml stainless autoclave equipped with an electromag- rate of 10 °C min1. The ion trap detector had a mass range of
netic-driven stirrer. For each run, 1.5 g catalyst, 40 ml water and 50–500 m/z with scan times of 1 s. The mass spectrometer ion
2.0 g phenol were loaded into the autoclave. After displacing air, source was 250 °C with a 70-eV ionization potential. The concen-
the H2 pressure was raised to 4.0 MPa and maintained with a back tration of Ni ions remaining in the filtered reaction solution was
pressure regulator. Continuous hydrogen flow of 200 mL/min was determined by ICP-AES (Inductively coupled plasma-atomic emis-
controlled by a mass flow-meter. Reactions were carried out at sion spectroscopy, IRIS Advantage 1000, Thermo Electron Corpora-
160–240 °C for 3 h. Reagents were vigorously stirred at 800 rpm. tion). CO, CO2, CH4 were analyzed by GC with a thermal
Tail gas samples were collected in vacuum plastic bags for subse- conductivity detector, TDX-01 (3 m  3 mm) column, and C2H4
quent off-line gas chromatography (GC) analysis. and C2H6 were analyzed by GC with a Flame Ionization Detector,
The recyclability of 10 wt.% Ni/HZSM-5 (Si/Al = 38) catalyst was Porapak-Q (3 m  3.175 mm).
tested in the 100-ml stainless autoclave with 2.0 g phenol and
40 ml water at 240 °C. Upgrading of bio-oil was also carried out
in the 100-ml stainless autoclave with 40 ml of raw bio-oil. 3. Result and discussion

3.1. Characterization of catalysts


2.3. Characterization of catalysts
The textural properties of the Ni-based catalysts with different
The BET (Brunauer-Emmett-Teller) surface area, average pore support materials are shown in Table 1. The Ni/HZSM-5 catalysts
diameter and pore volume of catalysts were determined by N2 iso- with different ratios of Si/Al had similar textural properties,
thermal (77 K) adsorption using the Micrometrics ASAP-2010 though the BET surface area, micropore area, micropore volume,
automated system. and pore size of Ni/HZSM-5(Si/Al = 38) catalyst were better than
NH3 temperature-programmed desorption (NH3-TPD) studies those of the Ni/HZSM-5(Si/Al = 50) catalyst. For Ni/c-Al2O3
were carried out in a quartz tube reactor with a thermal conductiv- catalyst, the BET surface area was lower than that of Ni/HZSM-5
ity detector (TCD). One-hundred mg of catalyst was pretreated in a catalysts, and the micropore area was very small, but pore size
flow of helium (30 ml/min) at 400 °C for 1 h, and after cooling to and total pore volume were higher than that of the Ni/HZSM-5
100 °C, ammonia adsorption was carried out. Subsequently, exces- catalysts, which means the HZSM-5-supported catalyst had a mes-
sive physisorbed ammonia was removed by purging with helium opore structure.
at 100 °C for 2 h. All NH3-TPD tests were carried out by increasing The overall acid amounts in three catalysts can be inferred from
the temperature from 100 to 650 °C at a rate of 10 °C/min and a he- the relative peak areas of the NH3 desorption curves. The NH3-TPD
lium flow rate of 30 ml/min. Thermogravimetry (TG) studies of profiles for various Ni-based catalysts are shown in Fig. 1. The
used catalysts were carried out under an air flow rate of 30 ml/ overall acid amounts decreased slightly with increasing Ni loading,
min with STA 409 PC thermal analyzer (NETZSCH, Germany) by and the acid amounts of Ni/HZSM-5 catalysts decreased slightly
using 10–15 mg samples and a 10 °C/min temperature increase. with the increase in the Si/Al ratio. Moreover, the strength of acid
Fresh and used catalysts were characterized by X-ray diffraction sites also decreased slightly with increasing Si/Al ratios. The acid
(XRD) (XPert Pro MPD with Cu Ka(k = 0.154) radiation, PANalytical, amounts of the 10 wt.% Ni/c-Al2O3 catalyst were very small com-
Netherlands). pared with those of the Ni/HZSM-5 catalysts, and its strength of
308 X. Zhang et al. / Bioresource Technology 127 (2013) 306–311

Table 1
Textural properties of the Ni-based catalysts.

Catalysts BET (m2/g) Micropore area (m2/g) Total pore volume (m3/g) Micropore volume (m3/g) Pore size (nm)a
10 wt.% Ni/HZSM-5 (Si/Al = 38) 290.4 203.7 0.161 0.095 2.2
10 wt.% Ni/HZSM-5 (Si/Al = 50) 227.0 163.9 0.115 0.076 2.0
10 wt.% Ni/c-Al2O3 135.3 2.41 0.269 – 7.9
a
Adsorption average pore diameter: calculated by 4 V/S (V: total pore volume; S: BET surface area).

Fig. 2. Product distribution for the hydrodeoxygenation of phenol over different


Fig. 1. NH3-TPD profiles of Ni-based catalysts. (a) 6 wt.% Ni/HZSM-5, Si/Al = 38; (b) catalysts. Reaction temperature: 240 °C; reaction time: 3 h; pressure: 4.0 MPa. Data
10 wt.% Ni/HZSM-5, Si/Al = 38; (c) 10 wt.% Ni/HZSM-5, Si/Al = 50; (d) 14 wt.% Ni/ are the average of results from two experiments, the experimental error is 62.3%.
HZSM-5, Si/Al = 38; (e) 10 wt.% Ni/c-Al2O3.
loadings, and the activity of the Ni/c-Al2O3 catalyst was the lowest.
The conversion of phenol HDO over Ni/c-Al2O3 was only 20.9%.
acid sites was also weak. Thus, the order of catalysts acidity for the One of the reasons may be related to the small BET surface area
same Ni loading was Ni/HZSM-5(Si/Al = 38) > Ni/HZSM-5(Si/ and the microspore area of catalyst Ni/c-Al2O3.
Al = 50) > Ni/c-Al2O3. The distribution of products resulting from HDO of phenol over
catalysts with different supports and Ni loading of 10 wt.% at
3.2. Catalytic activity 240 °C is shown in Fig. 2. The primary product was benzene for
all catalysts, and the selectivity for benzene was more than 60%.
The catalytic activity was evaluated by the phenol HDO reaction Cyclohexane, cyclohexene, cyclohexanol and methyl-cyclopentane
in water. Phenol was chosen as a model compound for bio-oil since were also observed. The selectivity for cyclohexanol decreased
phenol and its derivatives are the main components of bio-oil. gradually in the order of Ni/c-Al2O3, Ni/HZSM-5(Si/Al = 50), and
Moreover, phenol is considered to be a stable compound. Ni/HZSM-5(Si/Al = 38). This trend parallels that for the catalysts
acidity. In contrast, the selectivity for cyclohexane decreased in a
3.2.1. Effect of supporting material on catalytic activity and selectivity reverse order, which suggests that strong acidity favors formation
The conversion of phenol HDO obtained over different catalysts of cyclohexane via dehydration of cyclohexanol. The methyl-cyclo-
is presented in Table 2. The Ni/HZSM-5(Si/Al = 38) catalysts could pentane selectivity over 10 wt.% Ni/HZSM-5(Si/Al = 38) was higher
effectively convert phenol to hydrocarbons by HDO reaction in than that of the other two catalysts (Fig. 2). This outcome may also
water, especially at a Ni loading of 10 wt.%. The supporting mate- be related to the acidity of the supports because a strongly acidic
rial had a great effect on catalytic activity. The Ni/HZSM-5(Si/ catalyst favors the isomerization of intermediates formed during
Al = 38) was more active than the Ni/HZSM-5(Si/Al = 50) for all Ni the phenol HDO (Kazakov et al., 2011).

3.2.2. Temperature effect on catalytic activity and selectivity


Table 2
Aqueous-phase hydrodeoxygenation of phenol over Ni-containing catalysts. Changes in the conversion of phenol and product distribution
with reaction temperature over 10 wt.% Ni/HZSM-5 catalyst are
Catalysts Si/Al ratio (molar ratio) Conversion /%
presented in Fig. 3. At the investigated temperature, the catalyst
6 wt.% Ni/HZSM-5 38 56.3 (2.3) exhibited excellent activity for phenol conversion, and a tempera-
10 wt.% Ni/HZSM-5 38 91.8 (1.9)
ture increase resulted in a phenol conversion increase.
14 wt.% Ni/HZSM-5 38 85.7 (1.3)
6 wt.% Ni/HZSM-5 50 22.6 (2.1)
The possible pathway of phenol HDO has been reported
10 wt.% Ni/HZSM-5 50 41.9 (1.1) (Furimsky, 2000; Gandarias et al., 2008; Yang et al., 2009; Zhao
14 wt.% Ni/HZSM-5 50 33.7 (1.6) et al., 2009) and consists of two main routes: direct elimination
10 wt.% Ni/c-Al2O3 – 20.9 (2.7) of the hydroxyl group by hydrogenolysis, and indirect elimination
Reaction conditions: Temperature: 240 °C; reaction time: 3 h; H2 pressure: 4.0 MPa. of the hydroxyl group by thermal dehydration of a saturated cyclic
Data are the average of results from two experiments. Data listed in the brackets are alcohol formed by hydrogen addition to the aromatic ring. In this
the experimental error. study, the main products derived from phenol HDO in water were
X. Zhang et al. / Bioresource Technology 127 (2013) 306–311 309

Fig. 4. Thermogravimetry of fresh and used 10 wt.% Ni/HZSM-5 (Si/Al = 38)


Fig. 3. Products distribution and phenol conversion as a function of temperature. catalyst. (a) Fresh catalyst treated by dipping in the phenol solution; (b) catalyst
Data are the average of results from two experiments; the experimental error is used in aqueous phase HDO of phenol at 240 °C.
63.1%.

benzene, cyclohexane, cyclohexene, cyclohexanol and methyl- Table 3


cyclopentane. This observation suggests that both reaction routes Properties of raw and upgraded bio-oil.
might take place simultaneously during the process of phenol
Raw bio-oil Upgraded bio-oil
HDO in water. According to Fig. 3, hydrogenolysis of phenol to
benzene was more favorable at a higher temperature, and the indi- pH 2.27 (0.11) 4.07 (0.07)
KOH titration value (mol/L) 0.98 (0.08) 0.26 (0.08)
rect elimination of hydroxyl group via cyclic alcohol intermediate
Water content/wt.% 38.3 (0.32) 52.01 (0.43)
is the primary reaction route at a lower temperature. This conclu- Heat value/MJ kg1 13.79 (0.28) 14.32 (0.21)
sion is in a good agreement with those of Yang et al. (2009) and
Element compositions (wt.%)
Gandarias et al. (2008). C 53.59 (0.72) 54.33 (0.58)
H 6.28 (0.07) 7.01 (0.08)
3.2.3. Catalyst recyclability O 40.1 (1.12) 38.66 (0.66)
N 0.03 (0.01) –
The 10 wt.% Ni/HZSM-5(Si/Al = 38) catalyst could be used thrice
while retaining a high catalytic activity (phenol conversion: 91.8% Data are determined in duplicate. Data listed in the brackets are the experimental
(first), 90.3% (second) and 87.5% (third) respectively); however, a error.
significant drop in conversion of phenol (57.7%) was observed
when the catalyst was reused fourthly at the temperature of contained in upgraded bio-oil, residual polymers and cokes vs.
240 °C. Typical reasons for Ni based catalyst deactivation include carbon contained in the raw bio-oil was 95.6%. The raw bio-oil
sintering, coking, and leaching. In this study, the calcination and was obtained from fast pyrolysis of pine sawdust and its properties
reduction temperature of the catalysts was 450 °C and the reaction (pH value, water content and element composition) are presented
temperature was not in excess of 240 °C, hence sintering was unli- in Table 3.
kely. To better understand the catalyst deactivation mechanism, The FTIR spectra for the upgraded bio-oil and raw bio-oil sam-
TG and XRD analyses were performed on fresh and used catalysts. ples are presented in Fig. 5. The absorption bands centered at about
The TG curve of the used catalyst (10 wt.% Ni/HZSM-5(Si/Al = 38) 3400 cm1 are assigned to typical hydroxyl group absorption and
used in the aqueous phase HDO of phenol at 240 °C) is illustrated may originate from a combination and overlap of aliphatic and aro-
in Fig. 4. There is a weight loss of about 6 wt.% for the used catalyst, matic O–H as well as free and bound water contained in the bio-oil.
which may be attributed to the desorption of residual phenol ab- The intensity of the absorbance at this range intensified after hy-
sorbed on catalysts. Only about 2 wt.% coke formed on the catalyst dro-treating, implying that the amount of increased due to the for-
during the HDO reaction (Fig. 4). Thus coke formation may not be mation of water during the hydrotreatment process. The
the main reason for catalyst deactivation. absorption bands in the region of 3020–2940 cm1 are assigned
The X-ray diffraction patterns of reduced fresh catalyst and to carboxyl group absorption. The absorption bands became very
used catalyst are presented in Supplemental Fig. S1. The peaks at small after hydrotreating. This implies that the amount of carboxyl
about 51.9°, 44.6° are ascribed to the diffraction peaks of Ni metal. groups decreased due to the decarboxylation reactions during
The intensities of these peaks in used catalyst weaken, or even dis- hydrotreating of bio-oil. The absorption bands in the region of
appear, after repeated use due to leaching of Ni. The presence of 2230–2070 cm1 are assigned to the stretching vibrations of
89 mg/L of Ni in the filtered reaction solution also indicates that C„C groups. Absorbance in this region weakened after hydrotreat-
leaching of Ni occurred during the hydro-treating process. ing. The absorption bands in the region of 1730–1580 cm1 are as-
signed to the stretching vibrations of C@C groups in aromatics. The
3.3. Upgrading of bio-oil absorptions bands between 650–500 cm1 may be attributed to
the deformation vibrations of aromatic rings. Compared with the
The upgrading experiment of raw bio-oil was carried out under FTIR spectra of raw bio-oil, the absorbance of the absorption bands
the optimum conditions established for phenol (10 wt.% Ni/HZSM- at 640 and 505 cm1 intensified. This result indicates that the
5(Si/Al = 38) catalyst and 240 °C). For the upgrading of bio-oil, the amount of aromatic groups in the bio-oil increased rapidly after
carbon balance estimated by the sum of CO, CO2, CH4, C2H6 carbon hydrotreating over 10 wt.% Ni/HZSM-5 catalyst.
310 X. Zhang et al. / Bioresource Technology 127 (2013) 306–311

In order to investigate the changes for main components of raw


and upgraded bio-oil, GC–MS analysis were carried out. There are
hundreds of peaks in the GC–MS spectrum (Supplemental
Fig. S2) in bio-oil and the compounds listed in Table 4 are those
represented by peak areas larger than 1% of total. Esters such as
ethyl acetate, methyl propionate, formic acid 1-methylpropyl ester
and propanoic acid propyl ester increased after hydrotreating in
agreement with the FTIR results. Furfural was 5.95% in the raw
bio-oil, but disappeared after hydrotreating and some phenolic
compounds such as 2-methoxy-phenol and 1,2-benzenediol also
decreased after hydrotreating, possibly because they were con-
verted into coke-like polymers in the presence of a solid acid cat-
alyst (Zhao et al., 2009; Mohapatra et al., 1998). Organic acids
such as acetic and propanoic acids also decreased after
hydrotreating.
To evaluate the properties of the upgraded bio-oil, pH and water
content were quantified (Table 4). After hydrotreating over 10 wt.%
Ni/HZSM-5(Si/Al = 38) catalyst, the pH of the bio-oil increased
from 2.27 to 4.07 due to the decrease in organic acids. Xu et al.
Fig. 5. FTIR spectrum of upgraded and raw bio-oil samples. The condition of hydro- (2010) previously reported a pH increase only from 2.33 to 2.77
treatment for upgrading of bio-oil: catalyst: 10 wt.% Ni/HZSM-5 (Si/Al = 38);
after upgrading over a MoNi/c-Al2O3 catalyst. Water content in
temperature: 240 °C; reaction time: 3 h; H2 pressure: 4.0 MPa.
the bio-oil increased from 38.3% to 52.01% via hydrotreating. The
water could have been produced from the HDO reaction or from
undesirable repolymerisation reactions (Venderbosch et al., 2010;
Strengthening of the absorption peak centered at 1400 cm1 Mercader et al., 2011). The hydrogen content increased from 6.28
attributed to the stretching vibrations of alkanes, implies that the to 7.01 wt.% which revealed the hydrotreatment did happen during
alkane concentration increased after upgrading. The absorption the upgrading process. Correspondingly, the calorific value in-
peak centered at about 1230 cm1 is assigned to the stretching creased from 13.79 to 14.32 MJ/kg. The increase in the amount of
vibrations of C–O groups which are contained in esters and ethers. water in the upgraded bio-oil may be the reason for the only minor
Its absorbance intensified after hydrotreating. Li et al. (2011) and increase in the calorific value. The oxygen content decreased only
Xu et al. (2009) reported that esters could be formed from acetic slightly, possibly because the water was not removed prior to the
acid under high H2 pressure over Ni-based catalysts. In the present elemental analysis of raw and upgraded bio-oil.
study, esters could have been formed from carboxyl compounds About 36 wt.% residues were deposited on the surface of the
contained in the bio-oil. catalyst based on thermogravimetry (Supplemental Fig. S3). Thus,
it is possible that most of the activate sites of catalyst were covered
by coke-like residues during the hydro-treating of bio-oil. Thus,
Table 4
Main components of raw bio-oil and upgraded bio-oil as determined by GC–MS. improvements and optimization of catalysts and reaction condi-
tions to avoid coke formation are required.
Name Raw bio-oil/ Upgraded bio-oil/
Area% Area%
Propanoic acid,2-oxo-,ethyl ester 3.17 –
4. Conclusions
Acetaldehyde,hydroxyl- 2.33 –
2,3-Butanedione 1.59 –
3-Pentanone 3.97 1.68 Ni/HZSM-5(Si/Al = 38) catalysts showed higher activity than
Ethyl acetate 0.88 1.99 Ni/HZSM-5(Si/Al = 50) and Ni/c-Al2O3 catalysts for phenol hydro-
Methyl propionate – 6.51
deoxygenation at 240 °C. For the Ni/HZSM-5(Si/Al = 38) catalysts,
Acetic acid 9.41 4.88
2-Propanone,1-hydroxy- 15.67 5.21
the most suitable Ni loading was 10 wt.%. Upgrading of bio-oil in
2-Butanone,-hydroxy- – 5.77 the presence of the Ni/HZSM-5 catalyst increased the pH to 4.07,
Propanoic acid 2.39 0.83 the hydrogen content from 6.28 to 7.01 wt.%, and the gross calorific
Acetic acid,(acetyloxy)- 1.09 3.32 value from 13.79 to 14.32 MJ/kg. This work proved that the prop-
Propanal 2.82 –
erties of bio-oil, particularly the hydrogen content and the acidity
1-Butanol – 4.11
Formic acid,1-methylpropyl ester – 2.64 can be effectively improved though hydrotreating.
Furfural 5.95 –
3-Pentanol – 1.07
Propanoic acid,propyl ester – 1.28 Acknowledgements
2-Cyclopenten-1-one,2-methyl- 0.93 2.29
Butyrolactone 0.90 5.16
Cyclopentanone,2-methyl- 1.64 0.67 This work was supported by Grants from the National Natural
Phenol 1.47 1.03 Science Foundation of China (Project No. 51106167, 51076157),
Vinyl butyrate – 2.01 National science and technology Project founded by MOST of China
Phenol,3-methyl- 1.82 1.14 (Project No. 2011BAD22B07, 2012AA101808)
Phenol,2-methoxy- 4.01 2.2
2-Cyclopenten-1-one,3-ethyl-2- 0.2 1.37
hydro
Phenol,2-methoxy-4-methyl- 3.98 2.41 Appendix A. Supplementary data
1,2-Benzenediol 2.9 1.4
Phenol,4-ethyl-2-methoxy- 1.41 0.82
Supplementary data associated with this article can be found, in
4-Ethylcatechol 1.01 0.33
the online version, at http://dx.doi.org/10.1016/j.biortech.2012.07.
Compounds listed are those represented by more than 1% of the total peak area. 119.
X. Zhang et al. / Bioresource Technology 127 (2013) 306–311 311

References Mercader, F.D.M., Groeneveld, M.J., Kersten, S.R.A., Way, N.W.J., Schaverien, C.J.,
Hogendoorn, J.A., 2010b. Production of advanced biofuels: co-processing of
upgraded pyrolysis oil in standard refinery units. Applied Catalysis B:
Adjaye, J.D., Bakhshi, N.N., 1995. Production of hydrocarbons by catalytic upgrading
Environmental 96 (1–2), 57–66.
of a fast pyrolysis bio-oil. Part II: Comparative catalyst performance and
Mercader, F.D.M., Koehorst, P.J.J.J., Heeres, H.J., Kersten, S.R.A., Hogendoorn, J.A.,
reaction pathways. Fuel Processing Technology 45 (3), 185–202.
2011. Competition between hydrotreating and polymerization reactions during
Ardiyanti, A.R., Gutierrez, A., Honkela, M.L., Krause, A.O.I., Heeres, H.J., 2011.
pyrolysis oil hydrodeoxygenation. AIChE Journal 57 (11), 3160–3170.
Hydrotreatment of wood-based pyrolysis oil using zirconia-supported mono-
Mohapatra, D.K., Das, D., Nayak, P.L., Lenka, S., 1998. Polymers from renewable
and bimetallic (Pt, Pd, Rh) catalysts. Applied Catalysis, A: General 407 (1–2), 56–
resources: XX. Synthesis, structure, and thermal properties of semi-
66.
interpenetrating polymer networks based on cardanol–formaldehyde-
Bejblová, M., Zámostný, P., Červený, L., Čejka, J., 2005. Hydrodeoxygenation of
substituted aromatic compounds copolymerized resins and castor oil
benzophenone on Pd catalysts. Applied Catalysis, A: General 296 (2), 169–
polyurethanes. Journal of Applied Polymer Science 70 (5), 837–842.
175.
Oasmaa, A., Kuoppala, E., Ardiyanti, A., Venderbosch, R.H., Heeres, H.J., 2010.
Bunch, A.Y., Wang, X., Ozkan, U.S., 2007. Hydrodeoxygenation of benzofuran over
Characterization of hydrotreated fast pyrolysis liquids. Energy Fuels 24 (9),
sulfided and reduced Ni–Mo/c-Al2O3 catalysts: effect of H2S. Journal of
5264–5272.
Molecular Catalysis A: Chemical 270 (1–2), 264–272.
Peng, J., Chen, P., Lou, H., Zheng, X.M., 2009. Catalytic upgrading of bio-oil by HZSM-
Bui, V.N., Laurenti, D., Afanasiev, P., Geanter, C., 2011. Hydrodeoxygenation of
5 in sub- and super-critical ethanol. Bioresource Technology 100 (13), 3415–
guaiacol with CoMo catalysts. Part I: promoting effect of cobalt on HDO
3418.
selectivity and activity. Applied Catalysis, B: Environmental 101 (3–4), 239–
Scholze, B., Meier, D., 2001. Characterization of the water-insoluble fraction from
245.
pyrolysis oil (pyrolytic lignin). Part I. PY-GC/MS, FTIR, and functional groups.
Díaz, E., Mohedano, A.F., Calvo, L., Gilarranz, M.A., Casas, J.A., Rodríguez, J.J., 2007.
Journal of Analytical and Applied pyrolysis 60 (1), 41–54.
Hydrogenation of phenol in aqueous phase with palladium on activated carbon
Senol, O.I., Ryymin, E.M., Viljava, T.R., Krause, A.O.I., 2007. Effect of hydrogen
catalysts. Chemical Engineering Journal 131 (1–3), 65–71.
sulphide on the hydrodeoxygenation of aromatic and aliphatic oxygenates on
Fisk, C.A., Margan, T., Ji, Y., Crocker, M., Crofcheck, C., Lewis, S.A., 2009. Bio-oil
sulphided catalysts. Journal of Molecular Catalysis A 277 (1-2), 107–112.
upgrading over platinum catalysts using in situ generated hydrogen. Applied
Vitolo, S., Bresci, B., Seggiani, M., Gallo, M.G., 2001. Catalytic upgrading of pyrolytic
Catalysis, A: General 358 (2), 150–156.
oils over HZSM-5 zeolite: behaviour of the catalyst when used in repeated
Furimsky, E., 2000. Catalytic hydrodeoxygenation. Applied Catalysis, A: General 199
upgrading-regenerating cycles. Fuel 80 (1), 17–26.
(2), 147–190.
Venderbosch, R.H., Ardiyanti, A.R., Wildschut, J., Oasmaa, A., 2010. Stabilization of
Gandarias, I., Barrio, V.L., Requies, J., Arias, P.L., Cambra, J.F., Güemez, M.B., 2008.
biomass-derived pyrolysis oils. Journal of Chemical Technology and
From biomass to fuels: hydrotreating of oxygenated compounds. International
Biotechnology 85 (5), 674–686.
Journal of Hydrogen Energy 33 (13), 3485–3488.
Xu, Y., Wang, T., Ma, L., Zhang, Q., Wang, L., 2009. Upgrading of liquid fuel from the
Gong, F., Yang, Z., Hong, C., Huang, W., Ning, S., Zhang, Z., Xu, Y., Li, Q., 2011.
vacuum pyrolysis of biomass over the Mo–Ni/c-Al2O3 catalysts. Biomass
Selective conversion of bio-oil to light olefins: controlling catalytic cracking for
Bioenergy 33 (8), 1030–1036.
maximum olefins. Bioresource Technology 102 (19), 9247–9254.
Xu, Y., Wang, T., Ma, L., Zhang, Q., Liang, W., 2010. Upgrading of the liquid fuel from
Kazakov, M.O., Lavrenov, A.V., Danilova, I.G., Belskaya, O.B., Duplyakin, V.K., 2011.
fast pyrolysis of biomass over MoNi/c-Al2O3 catalysts. Applied Energy 87 (9),
Hydroisomerization of benzene-containing gasoline fractions on a Pt/SO2 4 – 2886–2891.
ZrO2–Al2O3 catalyst: II. Effect of chemical composition on acidic and
Yang, Y., Gilbert, A., Xu, C., 2009. Hydrodeoxygenation of bio-crude in supercritical
hydrogenating and the occurrence of model isomerization reactions. Kinetics
hexane with sulfided CoMo and CoMoP catalysts supported on MgO: a model
and Catalysis 52 (4), 573–578.
compound study using phenol. Applied Catalysis, A: General 360 (2), 242–249.
Laurent, M., Delmon, B., 1994. Influence of water in the deactivation of a sulfided
Zakzeski, J., Bruijnincx, P.C.A., Jongerius, A.L., Weckhyysen, B.M., 2010. The catalytic
NiMo/c-Al2O3 catalyst during hydrodeoxygenation. Journal of Catalysis 146 (1),
valorization of lignin for the production of renewable chemicals. Chemical
281–285, 288–291.
Reviews 110 (6), 3552–3599.
Li, W., Pan, C., Sheng, L., Liu, Z., Chen, P., Lou, H., Zheng, X., 2011. Upgrading of high-
Zhang, S., Yan, Y., Li, T., Ren, Z., 2005. Upgrading of liquid fuel from the pyrolysis of
boiling fraction of bio-oil in supercritical methanol. Bioresource Technology 102
biomass. Bioresource Technology 96 (5), 545–550.
(19), 9223–9228.
Zhao, C., Kou, Y., Angeliki, A.L., Li, X., Johannes, A.L., 2009. Highly selective catalytic
Mercader, F.D.M., Groeneveld, M.J., Kersten, S.R.A., Venderbosch, R.H., Hogendoorn,
conversion of phenolic bio-oil to alkane. Angewandte Chemie International
J.A., 2010a. Pyrolysis oil upgrading by high pressure thermal treatment. Fuel 89
Edition 48 (22), 3987–3990.
(10), 2829–2837.

Anda mungkin juga menyukai