Вы находитесь на странице: 1из 8

Bioresource Technology 134 (2013) 7380

Contents lists available at SciVerse ScienceDirect

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

Hydrodeoxygenation of lignin-derived phenolic compounds


to hydrocarbons over Ni/SiO2ZrO2 catalysts
Xinghua Zhang a,b, Qi Zhang a, Tiejun Wang a,, Longlong Ma a,, Yuxiao Yu a,b, Lungang Chen a
a
b

Key Laboratory of Renewable Energy and Gas Hydrate, Guangzhou Institute of Energy Conversion, Chinese Academy of Sciences, Guangzhou 510640, PR China
University of Chinese Academy of Science, Beijing 100049, PR China

h i g h l i g h t s

g r a p h i c a l a b s t r a c t

" SiO2ZrO2 exhibited amphoteric

SiO2ZrO2 supported Ni catalysts are attractive for hydrodeoxygenation of phenolic compounds. With
these catalysts, real phenolic compounds obtained from lignin degradation can be efciently converted
into hydrocarbons. This is meaningful to produce high-grade liquid fuels from lignin.

"
"
"

"

character and excellent textural


properties.
Ni/SiO2ZrO2 catalyst showed the
catalytic activity for phenolics HDO.
The bond of CO could be cleaved on
the surface of SiO2ZrO2 support.
Ni/SiO2ZrO2 catalyst exhibits
excellent recyclability and anticoking performance.
The lignin-derived phenols were
converted into hydrocarbons with
high yields.

OH

OCH3

Ni

OH

HO
HO

HO

OCH3

H H

CH3

O
OH

HO

H3C

lignin

Zr

Si
O

Zr
O

Si
O

O
R

OH

OCH3

H2

HDO

Depolymerization

OH

OH
O
OCH3
OH

OCH3
Lignin

a r t i c l e

i n f o

Article history:
Received 21 November 2012
Received in revised form 13 February 2013
Accepted 14 February 2013
Available online 22 February 2013
Keywords:
SiO2ZrO2
Phenolic compounds
Hydrodeoxygenation (HDO)
Hydrocarbon

Lignin-derived
phenolic compounds

Hydrocarbons

a b s t r a c t
Inexpensive non-sulded Ni-based catalysts were evaluated for hydrodeoxygenation (HDO) using guaiacol as model compound. SiO2ZrO2 (SZ), a complex oxide synthesized by precipitation method with different ratio of Si/Zr, was impregnated with Ni(NO3)26H2O and calcined at 500 C. Conversion rates and
product distribution for guaiacol HDO at 200340 C were determined. Guaiacol conversion reached
the maximum at 300 C in the presence of Ni/SZ-3. When HDO reaction was carried out with real lignin-derived phenolic compounds under the optimal conditions determined for guaiacol, the total yield
of hydrocarbons was 62.81%. These hydrocarbons were comprised of cyclohexane, alkyl-substituted
cyclohexane and alkyl-substituted benzene. They have high octane number, would be the most desirable
components for fungible liquid transportation fuel.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
The use of biomass is growing rapidly because of increasing
interests in renewable raw materials and zero emission (Zakzeski
et al., 2010). Typically, carbohydrate components could be converted into ethanol fuels (Aswathy et al., 2010; Singh et al.,

Corresponding authors. Tel.: +86 20 87057673; fax: +86 20 87057789.


E-mail addresses: wangtj@ms.giec.ac.cn (T. Wang), mall@ms.giec.ac.cn (L. Ma).
0960-8524/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.biortech.2013.02.039

2010), alkane fuels (Huber et al., 2004) and chemical intermediates


(Sahu and Dhepe, 2012) via diverse processes. In these processes,
lignin is removed as residue. Lignin can be broken down to phenolic compounds, such as alkyl-substituted phenols, guaiacols and
syringols by various approaches (Shen et al., 2010; Torr et al.,
2011). These phenolic compounds cant be used as transportation
fuels directly due to their high oxygen content, poor chemical stability and immiscibility with hydrocarbon fuels (Yang et al., 2009).
Consequently, for better uses of phenolic compounds derived from
lignin, process of hydrodeoxygenation (HDO) is needed in order to

74

X. Zhang et al. / Bioresource Technology 134 (2013) 7380

make these products more suitable as substitutes for the conventional fuels. In this work, it is desirable to use guaiacol for initial
screening since its structure is similar to the monomer of lignin
containing three types of CO bonds of CArOH, CArOCH3 and CArOCH3, then investigating the other model compounds and real lignin-derived phenolic compounds.
In the eld of HDO of oxygenated compounds, conventional
hydrodesulfurization (HDS) catalystssulded CoMo and NiMo
were explored by the majority of researchers (Bredenberg et al.,
1982; Bunch et al., 2007; Senol et al., 2007). For these HDS catalysts, in order to maintain the activity and stability, introduction
of extrinsic sulfur-containing compounds is necessary to keep the
sulfurization of these catalysts during the catalyst activation and
HDO processes, which inevitably brought about sulfur pollution
in the nal products (Bui et al., 2011a).
In order to replace sulded Mo-based catalysts and seek for
higher yield of alkanes products, noble metal-based catalysts have
also been studied extensively (Zhao et al., 2009; Bejblov et al.,
2005; Gutierrez et al., 2009). The noble metal catalysts show excellent catalytic performance in the HDO of phenolics, and the main
products are composed of saturated hydrocarbons. For example,
in a diluted phosphorous acid, Zhao et al. (2009) converted several
phenolic model compounds to alkanes with high selectivity over
carbon-supported palladium catalysts. However, employment of
the noble metal-based catalysts could signicantly raise cost. Thus,
development of an inexpensive non-sulded catalyst seems more
reasonable.
Support is another key factor determining the HDO activity of
catalysts. The pristine c-Al2O3 was previously used to hydrotreat
oxygenated compounds for producing hydrocarbons, but severe
carbon deposition was observed (Bui et al., 2011b) Moreover, alumina is known to be metastable in the presence of water, and will
partially transform into boehmite under hydrothermal conditions
(Laurent and Delmon, 1994). To overcome these aws, catalysts
supported on SiO2 (Zhao et al., 2011; Yakovlev et al., 2009), active
carbon (Yan et al., 2008), TiO2 (Bui et al., 2011b), ZrO2 (Bui et al.,
2011a,b; Kim et al., 2012), zeolites (Bejblov et al., 2005), and various mixed oxides (Zakzeski et al., 2010) have been explored in recent years. It was reported that ZrO2 allowed CoMoS catalysts to
reach relatively high catalytic activities because the ZrO2 support
appears to favor activation of O-compounds on surface (Bui et al.,
2011b; Yakovlev et al., 2009). In addition, the carbon deposition
on the ZrO2-supported catalyst was lower than that of the sulded
CoMo/Al2O3 catalyst (Gutierrez et al., 2009). However, the texture
of ZrO2 is very poor. This is unfavorable for the improvement of
catalytic performance. The addition of SiO2 into ZrO2 may overcome this disadvantage. The interaction of ZrO2 and SiO2 can lead
to some unique physical and chemical properties (Dang et al.,
1995).
Herein, we report the effectiveness of the non-noble and nonsulded catalyst for the conversion of lignin-derived phenolic compounds into hydrocarbons. The binary SiO2ZrO2 was synthesized
as a support, and then the novel catalysts were prepared using
the pristine transition metal Ni as active component. Our strategy
was to convert the model compounds and real lignin-derived phenolic compounds into hydrocarbons via one pot method.

2. Methods
2.1. Catalyst preparation
A series of SiO2ZrO2 supports with different molar ratio of Si/Zr
were prepared with the following method. Appropriate amounts of
Na2SiO39H2O and ZrOCl28H2O were dissolved in distilled water
respectively. With continuous stirring, solution of ammonium ni-

trate was added dropwise into the solution of Na2SiO3 until its
pH reached an approximate value of 8.0, by which the Si(OH)4 precipitate was prepared. According to the similar procedure, the
Zr(OH)4 precipitate was prepared using ammonia as precipitator.
These two kinds of precipitates were mixed and whisked. The
mixed precipitate was aged for 12 h at the temperature of 75 C.
Subsequently, the precipitate was ltered and washed with distilled water to remove chloride ions completely. Solid obtained
was dried overnight at 120 C and then calcined at 500 C for 5 h,
by which the complex oxide was prepared. These samples are designated as SZ-n (n is the molar ratio of Si/Zr). Pure ZrO2 was prepared using the same method.
An appropriate amount of nickel nitrate was dissolved in a predetermined volume of distilled water based on the amount of the
supports, followed by agitating and evaporating to dryness. The solid that remained was dried in air at 120 C overnight and then calcined at 500 C for 5 h. The Ni loading was 10 wt.% for all catalysts
tested in this work. These catalysts were designated as Ni/SZ-n.
The catalysts were crushed, sieved through 75125 lm mesh,
and reduced at 500 C in a ow of reducing gas (5% H2 + 95% N2)
before use.

2.2. Catalyst characterization


The BET surface, surface area of microporous, average pore
diameter and pore volume of catalysts were determined by N2 isothermal (196 C) adsorption on Micrometrics ASAP-2010 automated system.
Catalysts were characterized By X-ray diffraction (XRD) (XPert
Pro MPD with Cu Ka (k = 0.154) radiation, PANalytical,
Netherlands).
NH3-Temperature Programmed Desorption (NH3-TPD) and CO2Temperature Programmed Desorption (CO2-TPD) researches of the
catalyst were carried out in a quartz tube reactor with a thermal
conductivity detector (TCD). Similar procedure was reported in a
literature (Zhang et al., 2013).
FT-IR spectra of synthesized supports were taken on a TENSOR27 FT-IR spectrometer using KBr pellet technique.
Thermogravimetry (TG) studies of used catalysts were carried
out under an air ow rate of 30 ml min1 with STA 409 PC thermal
analyzer(NETZSCH, Germany) using 1015 mg sample and a
10 K min1 temperature increasing.

2.3. Preparation of lignin-derived phenolic compounds


Alkali lignin was purchased from SigmaAldrich. The used lignin has an ash content of 19.0 wt.%, and has the following elemental composition (wt.%, dry basis): 47.7% C, 4.9% H, 0.1% N, 3.9% S,
respectively. The degradation reaction of lignin was performed in
a closed system of a stirred autoclave with durations of 1 h and
at temperatures of 280 C. In this work, 5 wt.% ZnCl2 was used as
the additives, and methanol was used as solvent. 1.0 MPa H2 was
injected into the reactor at room temperature. After the reaction,
the product mixtures were ltered. Char and reactor were washed
with acetone. Filtrate and the washed acetone were mixed, and
then acidied with 2 M HCl to a pH of 2. The water was removed
with anhydrous magnesium sulfate. The acidied liquid was ltered again, and then followed by evaporation using a rotary evaporator under reduced pressure. The obtained products contained
two immiscible phases referred to as light fraction and heavy fraction, which could be automatically separated by decantation. The
light fraction, used as real lignin-derived phenolic compounds for
the HDO reaction, was diluted with acetone for GCMS analysis.

75

X. Zhang et al. / Bioresource Technology 134 (2013) 7380

2.4. Catalyst testing

moles of HDO products

The HDO catalytic reactions of phenolic compounds were carried out in a 500 ml stainless autoclave equipped with an electromagnetic driven stirrer. For each run, catalyst (1.5 g), reactants
(10.0 g) and solvent (dodecane, AR, 99% 100.0 g) were loaded in
the autoclave. After displacing air, the reactor was pressurized with
H2 to 5.0 MPa. The autoclave was heated to a desired reaction temperature, while the reagents were vigorously stirred at a rate of
800 rpm. Liquid samples were withdrawn from the reactor on line
for subsequent off-line analysis. Each experiment of HDO catalytic
reaction was repeated twice. Data are the average of results from
two runs.
The reactants, including phenol, guaiacol, cresol, eugenol, transanethole and vanillin are commercial agents (AR grades, P98%,
Sinopharm Chemical Reagent Co. Ltd.). The mixtures of model
compound were comprised of phenol (1.0 g), guaiacol (5.0 g), cresol (1.0 g), eugenol (1.0 g), trans-anethole (1.0 g) and vanillin
(1.0 g). The lignin-derived phenolic compounds were obtained
from the degradation of alkali lignin.
2.5. Products analysis
For the HDO reactions of model compound, quantitative analysis of liquid products were performed by gas chromatography (GC)
(SHIMADZU GC2010 with a ame ionization detector (FID) and a
DB-5 column, 30 m  0.25 mm  0.1 lm, N2 as carrier gas with
99.995% purity), with commercial standard reagents as external
standards. The vaporization temperature was 250 C, and the oven
temperature program ranged from 50 to 250 C at the rate of
5 C min1.
Conversions of model compounds and product selectivity were
calculated based on the formulas:

mreactantin  mreactantout
 100%
mreactantin
mC P
 100%
C p Selectivity% P
mC P

Conersion%

1
2

Cp represents the content of products obtained from the HDO


reaction.
Qualitative and quantitative analysis of the lignin-derived phenolic compounds was determined by a Shimadzu GCMS-QP2010
equipped with a column of DB-5 (30 m  0.2 mm  0.2 lm), using
benzyl alcohol as internal standard. The carrier gas was He with
99.9995% purity and the oven temperature program ranged from
40 C (holding for 2 min) to 250 C (holding for 5 min) at a rate
of 10 C min1. The ion trap detector had a mass range of 50
500 m/z with scan times of 1 s. The mass spectrometer ion source
was 250 C with a 70-eV ionization potential.
For the HDO process of real lignin-derived phenolic compounds,
the mass balance was estimated using the following equations:
moles of identified components in reactants
X weight for the constituent components of of idenfied components

molecular weight of constituent components

mass balance=%

X weight for the constituent components of HDO reaction


molecular weight of constituent components
4

moles of HDO products


 100
moles of idenfied components in reactants

3. Results and discussion


3.1. Characterization of supports and catalysts
Porous structure of the catalysts was exhibited in Table 1. Compared with ZrO2 supported catalyst, an apparent increase in BET
surface area, microporous surface area and total porous volume
was observed for the SZ supported catalysts, indicating that the
addition of SiO2 leads to conspicuous changes in its physical properties. In addition, the BET surface area and microporous surface
area of catalysts slightly increased with increasing of Si/Zr ratio.
The XRD patterns of different supports and catalysts are shown
in Supplemental Fig. S1. ZrO2 has two coexisting structurally stable
phase: monoclinic and tetragonal phases. The 2h of 24.1, 28.2,
31.4, and 34.1 are assigned to the m-ZrO2, and the peaks centred
at 35.2, 50.2, 60.3 are assigned to the t-ZrO2. SZ complex oxide,
being an amorphous material, shows no distinct XRD peaks apart
from a very broad weak feature between 18 and 37. The characteristic peaks of NiO can be seen clearly in the XRD patterns of different catalysts. The crystallite sizes of NiO calculated by Scherrer
equation were also listed in Table 1. The crystallite size of NiO supported on the SZ was smaller than that of NiO supported on the
ZrO2. This suggests that the SZ composite can promote the dispersion of NiO.
The FT-IR spectra of the supports, Ni/ZrO2 and Ni/SZ-3 are recorded in the spectral range of 4000400 cm1 (Supplemental
Fig. S2). For the supports, the spectra show a sharp peak at
1100 cm1 indicating SiOSi asymmetric stretching vibration,
which is found at 1050 cm1 for pure silica. The shift in stretching
frequency is due to deteriorating silica framework after insertion of
Zr atom (Tarafdar et al., 2005). The peak at 960 cm1 indicates the
presence of SiOZr bond in the composite materials in accordance
with previous studies (Tarafdar et al., 2005). It implies that the solid phase reaction occurred between Si(OH)4 and Zr(OH)4 in calcination process. In addition, the FT-IR spectra of Ni/ZrO2 and Ni/SZ3 are similar to those of the supports, implying that the supported
Ni have no inuence on the support.
ZrO2 is an amphoteric oxide. To test the effect of SiO2 on
amphoteric characteristics of ZrO2, the NH3-TPD and CO2-TPD
experiments for different support samples were carried out (Supplemental Fig. S3). The observed desorption peaks in the NH3TPD proles show the acidity of SZ composite. In addition, noticed
in the NH3-TPD proles is that the highest peaks of SZ-3 moves to
higher temperatures, suggesting that the appropriate ratio of Si/Zr
could strengthen the acidity of SZ composite. The CO2-TPD proles
suggest that the alkalescence of SZ composite was in rough agreement with that of ZrO2.

Table 1
Porous structure characterization of different catalysts.

Catalysts

SBET (m2 g1)

Smicro (m2 g1)

Vtotal (m3 g1)

DP* (nm)

NiO size (nm)

Ni/ZrO2
Ni/SZ-1
Ni/SZ-3
Ni/SZ-5

6.21
221.6
228.8
230.0

11.7
12.3
16.1

0.064
0.69
0.69
0.71

41.4
12.5
12.1
12.3

50.7
29.9
29.3
29.0

The average pore diameter.


The catalysts were in oxide state, and the average sizes of NiO were calculated by Scherrer equation.

76

X. Zhang et al. / Bioresource Technology 134 (2013) 7380

Table 2
Conversion of guaiacol over different catalysts.
Entry

Catalysts

Conversion [%]

1
2
3
4
5
6

ZrO2
Ni/ZrO2
SZ-3
Ni/SZ-1
Ni/SZ-3
Ni/SZ-5

4.3
43.8
9.2
86.2
100
81.5

Reaction condition: T = 300 C, PH2 = 5.0 MPa, t = 8 h.

3.2. Guaiacol HDO reaction


Conversions for HDO reaction of guaiacol over different catalysts at 300 C are shown in Table 2. The experimental results indicate that the activities of catalysts supported on SZ are obviously
higher than that of the catalyst supported on ZrO2 under the same
reaction conditions. There are two possible reasons accounting for
this. Firstly, the BET surface area and porous structure of these catalysts supported on SZ mixed oxides are obviously superior to that
of ZrO2, which are favorable for the improvement of catalytic activity. Secondly, the acidity of SZ oxides is more than that of pure ZrO2
as discussed in NH3-TPD experiments. Acidity of catalyst were related to hydrogenation, isomerization, dehydration and cracking
functionality (Rana et al., 2000). Apart from the catalytic cleavage
of CO bond, the acid sites of supports can also catalyze dehydration reaction during the HDO reaction, which could co-work well
together with metal-catalyzed hydrogenation, and promote HDO
reaction (Zhang et al., 2010).
The guaiacol HDO reaction had been studied over different catalyst in the past. For example, Bui et al. (2011b) reported the HDO
of guaiacol over sulded CoMo catalysts supported on c-Al2O3 and
ZrO2 under the conditions of 300 C and 4 MPa hydrogen pressure.
However, oxygenated compounds such as phenol, methyl-phenol
still occupied a certain proportion in the total products, and the
selectivity for HDO products (oxygen-free products) was not in excess of 30%, although obtained 100% guaiacol conversion. Among
these HDO products, the dominant compound was benzene. However, in our work, the selectivity for total of hydrocarbons (benzene, cyclohexane, methyl-cyclohexane, and toluene) is almost
100% for the Ni/SZ-3 catalyst. The dominant product is cyclohexane
(Fig. 1). This result has important signicance for the production of
bio-fuel from lignin-derived phenolic compounds because current
regulations are restricting aromatics content in fuels.
Supports have important effect on the product distribution.
Apart from cyclohexane, it could be found that methylcyclohexane
and toluene have also taken some proportions in the products for
Ni/ZrO2 catalyst, while cyclohexane was the dominant product
for Ni/SZ-3 catalyst. In addition, the product distributions also varied with different ratio of Si/Zr. Compared with Ni/SZ-3 and Ni/SZ5 catalysts, the selectivity for oxygenated compounds such as anisole, phenol and catechol over Ni/SZ-1 catalyst is relatively high,
suggesting the activity of Ni/SZ-1 catalyst for HDO reaction is inferior to the catalysts of Ni/SZ-3 and Ni/SZ-5.
To test the catalytic activity of supports, the guaiacol HDO
experiments over supports of ZrO2 and SZ-3 alone were also carried out. The results are in good accordance with literature reported by Bui et al. (2011b). The catalytic activities of these
supports are very low (Table 2), and the main products are oxygenated compounds, such as catechol, methylcatechol, cresol and phenol. Thereinto, the catechol accounts for most of proportion (Fig. 1).
In addition, it should be noted that benzene was detected in the
products obtained over SZ-3. It is speculated that the benzene
was originated from cracking of CO bond over the acid sites of
SZ-3. Strong acidity favors activation of CO bond (Fogassy et al.,

Fig. 1. Distribution of main products over different catalysts. Reaction condition:


Catalyst: Ni/SZ-3, T = 300 C, PH2 = 5.0 MPa, t = 8 h.

2011) and results in the formation of benzene via the sequential


CO bond cleavage. The principle of this process is similar to the
catalytic conversion of bio-oil over ZSM-5(Zhang et al., 2011).

3.3. Reaction pathways of guaiacol HDO


The variation of conversion and product selectivity with reaction time over Ni/SZ-3 catalyst were presented in Fig. 2. Guaiacol
conversion increased from 25% to 93% in the range of 0.55 h.
The selectivity of cyclohexane also increased with reaction time,
and in excess of 90% after 3 h. At the beginning of the reaction,
the selectivities of catechol, benzene and toluene were relatively
high, and then decrease gradually with reaction time. The selectivity of phenol increased slightly at rst and then reduced with reaction time lengthened.
Methanol appeared at the initial stages of the guaiacol HDO
reaction, and then disappeared after 3 h. This result is different
from the reported guaiacol HDO reaction over CoMo/ZrO2 catalyst
in a continuous xed bed reactor, in which the methanol was observed in nal liquid product (Bui et al., 2011a,b). The reason of this
phenomenon is that methanol could be converted into CH4 and
H2O due to the methanol hydrogenolysis under the conditions of
hydrogen and hydrogenation catalyst. Similar reaction had been
reported, which showed that hydrogenation catalyst could promote the hydrogenolysis of methanol for the formation of CH4
and H2O (Klicpera and Zdrazil, 2002).

77

X. Zhang et al. / Bioresource Technology 134 (2013) 7380

Fig. 2. Conversion and product selectivity for guaiacol HDO at different reaction
time. Reaction condition: Catalyst: Ni/SZ-3, T = 300 C, PH2 = 5.0 MPa.

Several catalytic reaction pathways for guaiacol HDO reaction


over different catalysts have been proposed (Bui et al., 2011a; Gutierrez et al., 2009; Zhao et al., 2011; Lin et al., 2011). Different catalysts lead to different reaction pathways and product selectivity.
For the HDO reaction of guaiacol, the rst stage of reaction is
demethylation and demethoxylation over sulded CoMo catalysts
and the main product is benzene (Zhao et al., 2011). Over noble
catalysts, the rst stage of reaction is hydrogenation of aromatic
ring, and the main product is cyclohexane (Gutierrez et al., 2009;
Zhao et al., 2011). Considering the published literatures and the results of this work on the product selectivity at different reaction
time, tentative reaction pathways for guaiacol HDO are presented
in Supplemental Fig. S4.
The bond energies of CO bond in guaiacol molecule decreased
in the order: CArOH (phenolic hydroxyl, 414 kJ mol1) > CArOCH3
(methoxy group, 356 kJ mol1) > CArOCH3 (methyl, 247 kJ mol1)
(Bui et al., 2011a). In theory, the guaiacol HDO reaction should follow the demethylation route (route of (c) in Supplemental Fig. S4).
The rst reaction stage would be preceded by hydrogenolysis of
the CO bond to form catechol, followed by the elimination of
the hydroxyl groups to produce phenol and then cyclohexane. This
reaction sequence suggests that catechol and phenol would be the
primary reaction intermediates. This deduction coincided with the
fact that catechol and phenol were detected at the rst stage of
reaction.
It could be speculated that phenol could also be formed directly
through a preliminary hemolytic dissociation of the activated CAr
OCH3 giving radicals, which could be readily hydrogenated under
H2 pressure to form phenol and methanol. This is the reaction
route of demethoxylation (route of (b) in Supplemental Fig. S4).
The observed methanol is a strong evidence for the
demethoxylation.

Anisole was rarely mentioned in literatures as a reaction intermediate or product (Bui et al., 2011a,b; Zhao et al., 2009). In this
case, anisole was detected in reaction products. It could be speculated that anisole was formed from guaiacol via hydrogenolysis of
the CArOH bond. Theoretically, the CArOH could be activated easily by the active sites of catalyst surface due to the minimal steric
hindrance effect. However, the bond energy of CArOH is the strongest among the three types of CO bonds of guaiacol molecule.
Hence, the dehydroxylation route (route of (a) in Supplemental
Fig. S4).
It was reported that the oxygenated compounds containing polar CO bond would be adsorbed on the acid sites much more
strongly than less or no polar hydrocarbons (Fogassy et al.,
2011). For the guaiacol molecule, the CArOCH3 bond was activated
on the acidic sites of support surface, which cleaved heterolytically,
then the positively charged methyl group could be transferred to
aromatic ring, and the methyl-catechol formed (route of (d) in Supplemental Fig. S4) (Bui et al., 2011a). The methylcatechol was detected in the products of guaiacol HDO over support of SZ-3.
Methylcatechol could be transformed into toluene via two dehydroxylation reaction successively. And the toluene could be further
hydrogenated to form methylcyclohexane.
Summarizing the above discussion, it could be inferred that
dehydroxylation, demethoxylation, demethylation and transmethylation reaction might be parallel competing reactions at the rst
reaction step during the course of guaiacol HDO. In addition, the
selectivities of all detected intermediates were relatively low; none
of them was in excess of 10%. This result suggests that the rst reaction step is the rate-determining step for HDO reaction. The subsequent hydrogenation and dehydration of intermediates are faster,
and these intermediates would be converted into products quickly.
3.4. Effects of temperature, catalyst reusabitity on HDO of guaiacol
Reaction temperature has a pronounced effect on conversion and
product distributions (Table 3). In the investigated temperature
range, the conversion of guaiacol varied obviously with reaction
temperature over Ni/SZ-3 catalyst. However, despite the same guaiacol conversions were obtained at 300 and 340 C respectively, the
product distributions are clearly different. The selectivity for cyclohexane at 340 C is lower compared with the result obtained at
300 C, and the selectivity for benzene and toluene is relatively high
at 340 C. Moreover, bicyclic compounds appeared in the products
at 340 C, although minor. This implies that the intermolecular polymerization occurred at higher temperature. Phenolic mixtures tend
to polymerize (Zhao et al., 2009). The formed polymers might deposit on catalyst surface and cause deactivation.
There is only a slight weight loss in the TG curve of the Ni/SZ-3
catalyst used at 300 C (Supplemental Fig. S5). However, there is

Table 3
Conversion of guaiacol and product selectivity at different temperature.
Catalyst

Ni/SZ-3

Temperature [C]
Conversion [%]

200
14.6

250
67.3

0.7
21.3
3.2
7.7
1.1
1.8
27.5
36.7

1.1
93.3
1.7
0.9
0.6
0.7
1.7

Product distribution [%]


Benzene
Cyclohexane
Cyclohexene
Methyl-cyclohexane
Toluene
Anisole
Phenol
Catechol
Phenyl-anisole

Reaction condition: t = 8 h, P = 5.0 MPa.

300
100
0.2
96.8
1.6
1.1
0.3

340
100
6.3
60.2

1.8
0.5
0.9
10.3
18.4
1.6

78

X. Zhang et al. / Bioresource Technology 134 (2013) 7380

Table 4
The main components of lignin-derived phenolic compounds and their HDO products.*
Lignin-derived phenolic compounds

HDO products

Components

Content /%

Components

Yield/%

Cyclohexanol
Methylcyclopentanone
Phenol
Trimethylbenzene
Methylphenol (o,m,p)
Guaiacol
Dimethylphenol
Tetramethylbenzene
Ethylphenol
Methyl-guaiacol
Trimethylphenol
Isopropylphenol
Ethylmethylphenol
Trimethylanisole
Tetramethylphenol
5-Isopropyl-2-methylphenol
Propylanisole
4-Propyl-2-methylphenol
Eugenol
2-(1,1-Dimethylethyl)-4-methyl-phenol
Trimethylnaphthalene
Phenanthrene and its derivatives
Total

0.94
0.47
1.20
0.12
4.48
17.50
11.3
0.42
0.23
1.97
9.58
0.47
3.97
2.02
1.76
1.32
6.44
3.42
0.38
2.20
0.41
4.57
75.17

Total of cyclic hydrocarbons


Cyclic alkanes
Cyclohexane
Methylcyclohexane
Dimethylcyclohexane
Trimethylcyclohexane
Propylcyclohexane
1-Ethyl-3-methylcyclohexane
1-propyl-3-methylcyclohexane
1-(1,1-Dimethylethyl)-3-methylcyclohexane
1-Isopropyl-3-methylcyclohexane
Aromatic hydrocarbons
Toluene
Dimethylbenzene
Trimethylbenzene
2,4,5,7-Tetramethylphenanthrene
Phenolic compounds
Trimethylphenol
Tetramethylphenol

62.81
54.99
17.09
7.13
11.65
2.31
6.38
4.33
3.36
1.68
1.06
7.82
0.53
1.26
4.94
1.09
8.23
5.60
2.63

Reaction condition: Ni/SZ-3 catalyst, T = 300 C, PH2 = 5.0 MPa, t = 8 h.


Determined by GSMS using benzyl alcohol as internal standard.
The yields of HDO products were estimated by the mass of HDO products divided by mass of the identied lignin-derived phenolic compounds.

more than 14% weight loss for the catalyst used at 340 C. The weight
loss (approximately 5 wt.%) below 250 C may be attributed to the
removal of water and residual organics absorbed in the catalyst.
The weight loss (about 6.2 wt.%) between 250 and 360 C may be
attributed to the combustion/decomposition of residual polymer
deposited on the catalyst surface. The weight loss (about 3 wt.%)
above 360 C may be attributed to the combustion of cokes deposited
on catalyst surface. The weight loss for the catalyst used at 340 C
was higher than that of catalyst used at 300 C, which strongly suggests that higher reaction temperature could result in the formation
of polymers and cokes. Though guaiacol was converted completely at
340 C, a large part of guaiacol was converted into polymers, which
were the precursors of cokes and carbon deposit.
To investigate the recyclability of catalyst, a batch of Ni/SZ-3
was used repeatedly for guaiacol HDO at 300C and under initial
pressure of H2 5.0 MPa. The guaiacol conversions for four runs
were 100%, 91.1%, 89.8%, and 86.5%, respectively. A signicant drop
in conversion of guaiacol was observed after the catalyst was reused only once. And then the conversions decreased slightly in
the following two runs.
The repeatedly used catalyst was also analyzed by TG method.
Compared the TG curve with the catalyst used only once, the
weight loss increased inconspicuously while the catalyst was reused four times (Supplemental Fig. S5). This result may strongly
suggest that the catalyst show excellent resistance to coking formation. SZ mixed oxide exhibits excellent amphoteric character
as discussed previously. The amphoteric character of support could
inhibit coke formation (Bui et al., 2011b; Klicpera and Zdrazil,
2002). This conclusion was further conrmed by Yang et al.
(2009) through phenol HDO over CoMo catalysts supported on basic materials such as MgO, in which the reaction temperature is
relatively high (380450 C).

were carried out over Ni/SZ-3 catalyst at 300 C. The results were
presented in Supplemental Table S1. Phenol, o-cresol, eugenol
and trans-anethole were completely converted. In the products
of phenol and cresol HDO reaction, the selectivities of cyclohexane
and methyl-cyclohexane were 97.8%, 92.5%, respectively. For the
HDO reaction of eugenol and trans-anethole, propylcyclohexane
was the dominant product. In addition, a spot of 1-methyl-2-propylcyclopentane and 1-methyl-3-propylcyclohexane were also detected in the HDO products. Vanillin was tested under the same
conditions. 83% conversion was achieved with a selectivity of
30.7% cyclohexane, 64.9% methylcyclohexane, and 4.4% toluene.
Partial of vanillin might be converted into guaiacol due to the
decarbonylation under cracking condition (Samolada et al., 2000),
which might be the origin of the cyclohexane.
The HDO of mixtures was also investigated over Ni/SZ-3 catalysts at 300 C. Most of the phenolic compounds were converted
into cycloalkanes. The main products were cyclohexane, methylcyclohexane and propyl-cyclohexane correspondingly. And the
isomerized products, such as dimethylcyclopentane and 1methyl-2-propylcyclopentane, account for only a few. These results were similar to the results in the HDO of single phenolic compound, correspondingly. Interactions between components in the
mixture were not observed.
Summarizing the HDO reactions for the tested phenolic compounds, we found that these model compounds could be effectively converted into cyclic alkanes with high yields, which
suggests that the catalytic conversion routine of phenolic compounds HDO would be highly atom economic and energy efcient,
and also suggest that the Ni/SZ-3 catalyst can be applied widely for
diverse phenolic compounds conversion.

3.6. HDO reaction of real phenolic compounds derived from lignin


degradation
3.5. HDO reaction of other phenolic model compounds
Apart from guaiacol, HDO reactions for more phenolic compounds (phenol, o-cresol, eugenol, trans-anethole and vanillin)

The real phenolic compounds derived from lignin degradation


were complex mixtures. The contents of main components listed
in Table 4 were determined according to the results of GCMS

X. Zhang et al. / Bioresource Technology 134 (2013) 7380

using benzyl alcohol as the internal standard. Most of the detected


components are phenolic compounds with carbon atom number of
C6C10. Of these lignin-derived phenolic compounds, guaiacol was
the most abundant component with content of 17.5%. All of the
poly-alkyl-substituted phenolic compounds, such as dimethylphenol, trimethylphenol, 2-methyl-5-propenylphenol, account for
33.5% of total real phenolic compounds.
The HDO reactions for these real phenolic compounds were carried out over Ni/SZ-3 catalyst at 300 C. The yields of HDO products
were presented in Table 4. The yield for total of cyclic hydrocarbons was 62.81%, in which the yield of aromatic hydrocarbons
and cyclic alkanes were 7.82% and 54.99%, respectively. Compared
with the results of model compounds HDO reaction, some trimethylphenol and tetramethylphenol were detected in the real phenolic
compounds HDO products. The steric hindrance effect, which is related to the bulkiness of the alkyl-substituent and the number of
alkyl-substituents located on the aromatic ring, may be one of
the key factors that account for the inferior performance of hydrodeoxygenation and hydrogenation of the poly-substituted phenols
(Solladi-Cavallo et al., 2007).
The obtained oxygen-free products were mainly composed of
cyclohexane, methylcyclohexane, dimethylcyclohexane, propylcyclohexane, toluene, xylene, and other alkyl-substituted benzene (Table 4). The octane numbers of these oxygen-free products are quite
high. For example, the octane number (RON-denition) of cyclohexane, toluene and xylene is 97, 116, and 117, respectively. Moreover,
these oxygen-free products have vapor pressures and carbon atom
number within the range of gasoline. Thus, they would be the most
desirable components for a fungible liquid transportation fuel.
It was found that about 86% of the identied components in the
real phenolic compounds were converted to cyclic hydrocarbons.
However, it should be noted that the effects of the unidentied
compounds in the real phenolic compounds, which might be converted to cyclic hydrocarbons in the HDO process were not taken
into consideration. For mass balance, the mass loss may be mainly
attributed to two reasons: (i) the formation of coke and polymerlikes residues; (ii) the products contained in gas phase due to the
evaporation of volatile components such as cyclohexane.
4. Conclusions
SZ complex oxide was synthesized, and the novel Ni/SZ catalysts were prepared by impregnation method. The Ni/SZ-3 catalyst
exhibited excellent performance for HDO reaction of phenolic compounds at 300 C and 5.0 MPa H2 pressure. The model phenolic
compounds were efciently converted into hydrocarbons with
conversion of 100% and selectivity of above 98%. SZ oxide possesses
amphoteric character, which promotes the excellent recyclability
and anti-coking performance of Ni/SZ-3 catalyst in the HDO reaction of model phenolic compounds. Signicantly, the real ligninderived phenolic compounds were effectively converted into
hydrocarbons with alkane yield of 54.99% and aromatic hydrocarbons yield of 7.82%.
Acknowledgements
This work was supported by grants from the National Natural Science Foundation of China (Project No. 51106167, 51161140331), National Basic Research Program of China (973 program, Project No.
2012CB215304), National science and technology Project founded
by MOST of China (Project No. 2012AA101808).
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.biortech.2013.02.
039.

79

Reference
Aswathy, U.S., Sukumaran, R.K., Devi, G.L., Rajasree, K.P., Singhania, R.R., Pandey, A.,
2010. Bio-ethanol from water hyacinth biomass: an evaluation of enzymatic
saccharication strategy. Bioresource Technology 101 (3), 925930.
erveny, L., C
ejka, J., 2005. Hydrodeoxygenation of
Bejblov, M., Zmostny, P., C
benzophenone on Pd catalysts. Applied Catalysis, A: General 296 (2), 169175.
Bredenberg, J.B.S., Huuska, M., Rty, J., Korpio, M., 1982. Hydrogenolysis and
hydrocracking of the carbon-oxygen bond: I. Hydrocracking of some simple
aromatic O-compounds. Journal of Catalysis 77 (1), 242247.
Bui, V.N., Laurenti, D., Afanasiev, P., Geanter, C., 2011a. Hydrodeoxygenation of
guaiacol with CoMo catalysts. Part I: Promoting effect of cobalt on HDO
selectivity and activity. Applied Catalysis, B: Environmental 101 (34), 239
245.
Bui, V.N., Laurenti, D., Delichre, P., 2011b. Hydrodeoxygenation of guaiacol Part II:
Support effect for CoMoS catalysts on HDO activity and selectivity. Applied
Catalysis, B: Environmental 101 (34), 246255.
Bunch, A.Y., Wang, X., Ozkan, U.S., 2007. Hydrodeoxygenation of benzofuran over
sulded and reduced NiMo/c-Al2O3 catalysts: effect of H2S. Journal of
Molecular Catalysis A: Chemical 270 (12), 264272.
Dang, Z., Anderson, B.G., Amenomiya, Y., Morrow, B.A., 1995. Silica-supported
zirconia.1.Characterization by infrared spectroscopy, temperature-programmed
desorption, and X-ray diffraction. The Journal of Physical Chemistry 99 (39),
1443714443.
Fogassy, G., Thegarid, N., Schuurman, Y., Mirodatos, C., 2011. From biomass to biogasoline by FCC co-processing: effect of feed composition and catalyst structure
on product quality. Energy Environmental Science 4 (12), 50685076.
Gutierrez, A., Kaila, R.K., Honkela, M.L., Slioor, R., Krause, A.O.I., 2009.
Hydrodeoxygenation of guaiacol on noble metal catalysts. Catalysis Today
147 (34), 239246.
Huber, G.W., Cortright, R.D., Dumesic, J.A., 2004. Renewable alkanes by aqueousphase reforming of biomass-derived oxygenates. Angewandte Chemie
International Edition 43 (12), 15491551.
Kim, M., DiMaggio, C., Salley, S.O., Ng, K.Y.S., 2012. A new generation of zirconia
supported metal oxide catalysts for converting low grade renewable feedstocks
to biodiesel. Bioresource Technology 118, 3742.
Klicpera, T., Zdrazil, M., 2002. Preparation of high-activity MgO-supported CoMo
and NiMo sulde hydrodesulfurization catalysts. Journal of Catalysis 206 (2),
314320.
Laurent, M., Delmon, B., 1994. Inuence of water in the deactivation of a sulded
NiMo/c-Al2O3 catalyst during hydrodeoxygenation. Journal of Catalysis 146 (1),
281285, 288291.
Lin, Y.C., Li, C.L., Wan, H.P., Lee, H.T., Liu, C.F., 2011. Catalytic hydrodeoxygenation of
guaiacol on Rh-based and sulded CoMo and NiMo catalysts. Energy & Fuels 25
(3), 890896.
Rana, M.S., Srinivas, B.N., Maity, S.K., Dha, G.M., Rao, T.S.R.P., 2000. Origin of cracking
functionality of sulded (Ni) CoMo/SiO2ZrO2 catalysts. Journal of Catalysis 195
(1), 3137.
Sahu, R., Dhepe, P.L., 2012. A one-pot method for the selective conversion of
hemicellulose from crop waste into C5 sugars and furfural by using solid acid
catalysts. ChemSusChem 5 (4), 751761.
Samolada, M.C., Papafotica, A., Vasalos, I.A., 2000. Catalyst evaluation for catalytic
biomass pyrolysis. Energy & Fuels 14 (6), 11611167.
Senol, O.I., Ryymin, E.M., Viljava, T.R., Krause, A.O.I., 2007. Effect of hydrogen
sulphide on the hydrodeoxygenation of aromatic and aliphatic oxygenates on
sulphided catalysts. Journal of Molecular Catalysis A: Chemical 277 (12), 107
112.
Shen, D.K., Gu, S., Luo, K.H., Wang, S.R., Fang, M.X., 2010. The pyrolytic degradation
of wood-derived lignin from pulping process. Bioresource Technology 101 (15),
61366146.
Singh, A., Pant, D., Korres, N.E., Nizami, A.S., Prasad, S., Murphy, J.D., 2010. Key issues
in life cycle assessment of ethanol production from lignocellulosic biomass:
challenges and perspectives. Bioresource Technology 101 (13), 50035012.
Solladi-Cavallo, A., Baram, A., Choucair, E., Norouzi-Arasi, H., Schmitt, M., Garin, F.,
2007. Heterogeneous hydrogenation of substituted phenols over Al2O3 supported
ruthenium. Journal of Molecular Catalysis A: Chemical 273 (12), 9298.
Tarafdar, A., Panda, A.B., Pramanik, P., 2005. Synthesis of ZrO2SiO2 mesocomposite
with high ZrO2 content via a novel solgel method. Microporous and
Mesoporous Materials 84 (13), 223228.
Torr, K.M., van de Pas, D.J., Cazeils, E., Suckling, I.D., 2011. Mild hydrogenolysis of insitu and isolated Pinus radiata lignins. Bioresource Technology 102 (16), 7608
7611.
Yakovlev, V.A., Khromova, S.A., Sherstyuk, O.V., Dundich, V.O., Ermakov, D.Y.,
Novopashina, V.M., Lebedev, M.Y., Bulavchenko, O., Parmon, V.N., 2009.
Development of new catalytic systems for upgraded bio-fuels production
from bio-crude-oil and biodiesel. Catalysis Today 144 (34), 362366.
Yan, N., Zhao, C., Dyson, P.J., Wang, C., Liu, L., Kou, Y., 2008. Selective degradation of
wood Lignin over noble-metal catalysts in a two-step process. ChemSusChem 1
(7), 626629.
Yang, Y., Gilbert, A., Xu, C.B., 2009. Hydrodeoxygenation of bio-crude in supercritical
hexane with sulded CoMoP catalysts supported on MgO: a model compound
study using phenol. Applied Catalysis, A: General 360 (2), 242249.
Zakzeski, J., Bruijnincx, P.C.A., Jongerius, A.L., Weckhuysen, B.M., 2010. The catalytic
valorization of lignin for the production of renewable chemicals. Chemical
Reviews 110 (6), 35523592.

80

X. Zhang et al. / Bioresource Technology 134 (2013) 7380

Zhang, X., Wang, T., Ma, L., Wu, C., 2010. Aqueous-phase catalytic process for
production of pentane from furfural over nickel-based catalysts. Fuel 89 (10),
26972702.
Zhang, H., Cheng, Y., Vispute, T.P., Xiao, R., Huber, G.W., 2011. Catalytic conversion
of biomass-derived feedstocks into olens and aromatics with ZSM-5: the
hydrogen to carbon effective ratio. Energy and Environmental Science 4 (6),
22972307.

Zhang, X., Wang, T., Ma, L., Zhang, Q., Jiang, T., 2013. Hydrotreatment of bio-oil over
Ni-based catalyst. Bioresource Technology 127, 306311.
Zhao, C., Kou, Y., Angeliki, A.A., Li, X., Lercher, J.A., 2009. Highly selective catalytic
conversion of phenolic bio-oil to alkanes. Angewandte Chemie International
Edition 48 (22), 39873990.
Zhao, H.Y., Li, D., Bui, P., Oyama, S.T., 2011. Hydrodeoxygenation of guaiacol as
model compound for pyrolysis oil on transition metal phosphide
hydroprocessing catalysts. Applied Catalysis, A: General 391 (12), 305310.

Вам также может понравиться