You are on page 1of 9

Fuel 111 (2013) 709–717

Contents lists available at SciVerse ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Upgrading of bio-oil into advanced biofuels and chemicals. Part I.


Transformation of GC-detectable light species during the
hydrotreatment of bio-oil using Pd/C catalyst
Richard Gunawan, Xiang Li, Caroline Lievens, Mortaza Gholizadeh, Weerawut Chaiwat, Xun Hu,
Daniel Mourant, John Bromly, Chun-Zhu Li ⇑
Fuels and Energy Technology Institute, Curtin University, GPO Box U1987, Perth, WA 6845, Australia

h i g h l i g h t s

 A fast pyrolysis bio-oil was hydrotreated using Pd/C catalyst.


 Cyclopentanones are the main product of the hydrogenation of aldehydes.
 Propionic, butanoic and pentanoic acids are produced during hydrotreatment.
 Depolymerisation of the lignin-derived oligomer occurs during hydrotreatment.

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

Article history: Bio-oil cannot be directly used as biofuel mainly due to its abundant oxygen-containing functional
Received 24 December 2012 groups. Therefore, bio-oil must be upgraded to produce high quality biofuels. As the first part of this ser-
Received in revised form 15 March 2013 ies, this paper reports the transformation of GC-detectable light components during the hydrotreatment
Accepted 1 April 2013
of bio-oil using noble-metal catalyst Pd/C over a wide range of reaction temperature (150–300 °C). Our
Available online 16 April 2013
results show that aldehydes were easily hydrogenated, with cyclopentanone and 2-methyl cyclopenta-
none among the observed products. During hydrotreatment, carboxylic acids such as propionic, butanoic
Keywords:
and pentanoic acids were also observed in the liquid products. Furthermore, evolution of propyl guaiacol,
Bio-oil
Upgrading
ethyl guaiacol, methyl guaiacol, guaiacol and syringol confirms the depolymerisation of the lignin-
Hydrotreatment derived oligomers during the hydrotreatment of bio-oil. The hydrogenation of the phenolic compounds
Palladium catalyst also took place. In addition, the FT-IR analysis of the O-containing functional groups, especially carbonyl
Biofuel groups, confirms the trends found by GC–MS.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction or diesel) biofuels from bio-oil will be reported in this series of


papers.
Bio-oil from the fast pyrolysis of biomass can potentially con- Apart from the conventional industrial hydrotreatment cata-
tribute to the reduced dependence on non-renewable fossil-de- lysts such as sulphided nickel-molybdenum (NiMo) and/or
rived liquid fuels. Unfortunately, its direct application in internal cobalt-molybdenum (CoMo) catalysts supported on c-alumina, no-
combustion engines is difficult due to its high water content (e.g. ble-metal catalysts such as Ru/C and Pd/C [2,3] were also explored
20 wt%), high acidity (pH 2–3), and overall high oxygen content for the hydrotreatment of bio-oil and model compounds due to
(30–50 wt%) [1] in a variety of chemical functionalities, which their known [4] activity for hydrogenation and hydrodeoxygen-
leads to high thermal instability and tendency of polymerisation. ation. The hydrotreatment of bio-oil was recently investigated in
Therefore, in order for bio-oil to be used as a liquid transportation a batch reactor using a range of noble-metal catalysts (Ru, Rh, Pt,
fuel, it needs to be upgraded to produce advanced biofuels. One of and Pd) with various supports (C, TiO2, and Al2O3) [5]. It was shown
the promising technological routes for upgrading is the catalytic that both yields and the levels of deoxygenation were higher using
hydrodeoxygenation of bio-oil. Our recent efforts to investigate the noble metal catalysts, especially Ru/C and Pd/C, than using the
the production of advanced ‘‘drop-in’’ (miscible with petrol and/ classical industrial catalysts (sulphided NiMo/Al2O3 and CoMo/
Al2O3).
Unfortunately, information regarding chemical transformations
⇑ Corresponding author. Tel.: +61 8 9266 1131; fax: +61 8 9266 1138. of different functional groups during upgrading is limited, which is
E-mail address: chun-zhu.li@curtin.edu.au (C.-Z. Li).

0016-2361/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.fuel.2013.04.002
710 R. Gunawan et al. / Fuel 111 (2013) 709–717

Table 1 description of the reactor system and experimental procedure


Yields of gas + coke, total water and total organics as a function of bio-oil can be found elsewhere [7]. The metal catalyst Pd/C (containing
hydrotreatment temperature and time.
5 wt% active metal) and chemicals used as standards were ob-
Temperature (°C) Time (h) Yields (wt% of raw bio-oil) of tained from Sigma Aldrich. All chemicals were used without fur-
Gas + coke Total water Total organics ther purification. Hydrogen and nitrogen (high purity grade)
150 3 3 23 74
were obtained from BOC.
200 3 9 26 65
250 3 9 28 64 2.2. Hydrotreatment
300 3 11 30 59
250 1 7 30 64 The catalytic hydrotreatment experiments were performed in a
250 3 9 28 64
250 6 7 30 63
batch autoclave reactor system described in the previous work
250 12 9 29 62 [8,9]. In a typical experiment, 4.5 g Pd/C catalyst was reduced in
300 1 12 31 57
40 bar H2 for 2 h at 300 °C. After cooling down and venting of gases,
300 3 11 29 61 raw bio-oil (90 g) was transferred into the reactor using an exter-
300 6 10 28 63 nal vacuum pump. The reactor was flushed with N2 and subse-
300 12 15 27 59 quently pressurised with H2 (40 bar) at room temperature. The
reactor was then heated to the desired reaction temperature
(150–300 °C) under continuous agitation at a rate of 1000 rpm.
mainly due to the complexity of bio-oil. Although various model
After the system reached the desired temperature, the hydrogen
compounds were utilised to understand the reactivities of different
pressure was kept constant at 100 bar. This was done by monitor-
components, individual model compounds and their mixtures
ing the pressure inside the reactor and adding hydrogen when the
show different reaction routes and therefore products [6]. Various
pressure decreased.
functional groups in a bio-oil will react with each other during
After the pre-determined reaction time, the reactor was rapidly
hydrotreatment. Therefore, care must be taken to understand the
cooled to room temperature by a cooling water coil. The liquid
hydrotreatment behaviour of bio-oil based on model compounds.
products, consisting of aqueous and oil phases, and the solids (cat-
As the first part of this series, this study reports our investiga-
alyst and coke) were recovered from the reactor and were weighed.
tion on the transformation of light species during the catalytic
They were then separated using a centrifuge. The gases formed
hydrotreatment of raw bio-oil using noble metal catalyst Pd/C un-
were quantified by the weight difference of the reaction mixture
der a wide range of reaction conditions. In addition to the light
before and after experiment.
components identified by GC–MS, efforts were made to understand
In some repeated experiments, the amounts of coke were deter-
the changes in carbonyl functionalities following the deconvolu-
mined by the increases in the weight of solids. After an experiment,
tion of FT-IR spectra. The changes of the aromatic structure and
the reactor content was collected and the reactor was then rinsed
the coke forming propensity will be revealed in the second part
thoroughly with acetone. The acetone-washed liquid with sus-
of this series of papers. The third part will demonstrate the impor-
pended solids was filtered. The filtered solids were then combined
tance of heavy species of bio-oil during its hydrotreatment in a
with wet solids separated from the liquid products after centrifu-
continuous packed-bed catalytic reactor. The two-stage hydro-
gation. A mixture of solvent (chloroform/methanol 4:1) was used
treatment, the effects of bio-oil pretreatment, the performance of
to wash the organics in the solid during filtration until the filtrate
various catalysts, and the novel reactor design for the hydrotreat-
is colourless. The filter was then dried and weighed. The amount of
ment of bio-oil will be among the topics for the later parts of this
solids subtracted by the weight of the catalyst loaded into the reac-
series.
tor was determined as the amount of solids (coke) formed during
hydrotreatment.
2. Experimental
2.3. Analyses
2.1. Materials
The water contents of the aqueous and oil phase liquid products
The raw bio-oil was obtained from the fast pyrolysis of mallee were determined by means of Karl-Fischer titration (Mettler Tole-
woody biomass in a fluidised bed reactor at 500 °C. The detailed do KF-DL31). About 0.3 g of sample was injected into an isolated

Fig. 1. Effects of (a) temperature (holding time: 3 h) and (b) reaction time on the water contents in aqueous and oil phases after hydrotreatment.
R. Gunawan et al. / Fuel 111 (2013) 709–717 711

Fig. 2. Total ion chromatograms of raw bio-oil and liquid products after 3 h of hydrotreatment at 250 °C.

glass chamber containing hydranal composite 5 K, and was titrated The distribution of the carbonyl functional groups of raw bio-oil
with hydranal ketosolver. All measurements were performed in and the liquid products was determined by FT-IR. Details of the
triplicates. FT-IR method can be found elsewhere [10,11].
The GC–MS analysis of the liquid products was performed using
an Agilent 5973 MSD attached to an Agilent 6890 GC, equipped 3. Results and discussion
with a 30 m  0.25 mm  0.25 lm HP-Innowax column (cross-
linked polyethylene glycol). 1 lL of methanol solutions containing 3.1. Phase distribution, yields and water contents
7–8 wt% of sample were injected into the injection port set at
250 °C with a split ratio of 50:1. The oven temperature was kept Almost all hydrotreatment experiments resulted in two liquid
at 40 °C for 3 min then heated up to 260 °C at a rate of 10 °C/min phases, i.e. a yellowish aqueous phase (40 wt% of the liquid prod-
and then held at 260 °C for 10 min. uct) and a brown oil phase (60 wt% of the liquid product) with a
712 R. Gunawan et al. / Fuel 111 (2013) 709–717

Table 2
Identification of the compounds labelled in Fig. 2.

Peak no Compound Formula Retention time (min)


1 Water H2O 3.37
2 Cyclopentanone, 2-methyl C6H10O 5.73
3 Methoxy methanol C2H6O2 6.05
4 2-Butanone, 3-hydroxy C4H8O2 7.29
5 Hydroxyacetone C3H6O2 7.53
6 Hydroxyacetaldehyde C2H4O2 7.77
7 Acetic acid C2H4O2 9.75
8 Furfural C5H4O2 9.92
9 2,5-Hexanedione C6H10O2 10.44
10 Propionic acid C3H6O2 10.91
11 1,2-Ethanediol C2H6O2 11.89
12 Butyrolactone C4H6O2 11.94
13 Butanoic acid C4H8O2 12.01
14 Pentanoic acid C5H10O2 13.28
15 2(5H)-furanone C4H4O2 13.40
16 2-Cyclopenten-1-one, 2-hydroxy C5H6O2 13.57
17 2-Cyclopenten-1-one, 2-hydroxy-3-methyl C6H8O2 14.21
18 Guaiacol C7H8O2 14.57
19 Guaiacol, 4-methyl C8H10O2 15.57
20 Phenol C6H6O 16.07
21 Guaicol, 4-ethyl C9H12O2 16.32
22 Guaiacol, 4-propyl C10H14O2 17.09
23 Syringol C8H10O3 18.51
24 1,2,4-Trimethoxybenzene C9H12O3 19.27
25 C9H10O4 C9H10O4 19.77
26 C14H12O C14H12O 20.37
27 Methoxyeugenol C11H14O3 22.37
28 Syringaldehyde C9H10O4 23.91
29 Levoglucosan C6H10O5 26.94
30 3,5-Dimethoxy-4-hydroxycinnamalde C11H12O4 28.85

Fig. 3. Effects of (a) temperature (holding time: 3 h) and (b) reaction time on the yields of cyclopentanone during the hydrotreatment of bio-oil.

density higher than the aqueous phase. Exception was the hydro- bio-oil. Reaction time was clearly an important parameter affecting
treatment of bio-oil at 150 °C where only single dark-brown liquid the formation of reaction products. However, it was interesting to
(oil phase) was obtained. This phenomenon was also reported in note that significant amounts of water were produced during the
the literature [12] for the mild-hydrotreating of bio-oil at 175 °C. first hour of hydrotreatment but with little additional formation
During the hydrotreatment of bio-oil, multiple reactions oc- even when the reaction time was extended up to 12 h. The coke
curred, including hydrogenation, hydrogenolysis, decarboxylation, formation was also determined for experiments at 250 °C and
decarbonylation, hydocracking and polymerisation leading to the 300 °C (both after 3 h). They were 1.20 wt% and 1.22 wt%, respec-
formation of coke. Clearly, through all of these reactions, gases tively. This is in agreement with the literature [5] that Pd/C gave
(CO, CO2, CH4, ethylene, ethane, propylene and propane), water, the least amount of coke among several noble metal catalysts
and solids (heavy polymer and coke) were produced [13]. Table 1 tested for the mild hydrotreating of bio-oil.
details the yields of gas + coke, total water and total organics as a Fig. 1a demonstrates that the water content in aqueous phase
function of reaction temperature and time. As expected, the higher increased while that in the oil phase decreased with increasing
the reaction temperature, the more gas and water were produced hydrotreatment temperature. The decreases in the water content
and the less organics produced from the hydrotreatment of of the oil phase indicate that more hydrophobic/non-polar
R. Gunawan et al. / Fuel 111 (2013) 709–717 713

Fig. 4. Effects of (a) temperature (holding time: 3 h) and (b) reaction time on the concentrations of 2-methyl cyclopentanone during the hydrotreatment of bio-oil.

Fig. 5. Effects of (a) temperature (holding time: 3 h) and (b) reaction time on the concentration of acetic acid during the hydrotreatment of bio-oil.

in the aqueous phase and moved to the oil phase. Similar trends
can be seen for the effect of reaction time during hydrotreatment
at 250 °C (see Fig. 1b). It is however interesting to see that the
water content in the oil phase decreased initially during the hydro-
treatment of bio-oil at 300 °C. This may imply that some hydrolysis
reactions could have taken place at this higher temperature of
300 °C.

3.2. Aldehydes and ketones

The light organics in bio-oil and upgraded products were ana-


lysed using GC–MS. The total ion chromatograms for typical aque-
ous and oil phases after hydrotreatment are compared with that of
the raw bio-oil in Fig. 2. The compounds identified in Fig. 2 are
listed in Table 2.
Among different functional groups products in bio-oil, alde-
hydes were the easiest to be hydrogenated to produce alcohols.
As is shown in Fig. 2, all aldehydes identified in raw bio-oil disap-
peared from the liquid products after hydrotreatment for 3 h at
Fig. 6. Yields of C3–C5 carboxylic acids as a function of reaction temperature during 250 °C. Hydroxyacetaldehyde and 3,5-dimethoxy-4-hydroxycin-
the hydrotreatment of bio-oil. namalde completely disappeared in the liquid products after
hydrotreatment even at temperatures as low as 150 °C.
compounds as well as water were produced during hydrotreat- Our data also show that 70.4% furfural was converted after 3 h
ment, forcing water to move to the aqueous phase. It is important hydrotreatment at 150 °C, and 98.8% was converted after 1 h
to note that although the water content changed significantly, the hydrotreatment at 250 °C. Under more severe conditions, furfural
yield ratio between aqueous phase and oil phase remained almost completely disappeared. Furfural can be hydrogenated to produce
the same [14], indicating that the organics must have also formed cyclopentanone which can be subsequently hydrogenated to
714 R. Gunawan et al. / Fuel 111 (2013) 709–717

Fig. 7. Yields of C3–C5 carboxylic acids as a function of reaction time during the hydrotreatment of bio-oil at (a) 250 °C and (b) 300 °C.

about 1 h of hydrotreatment at 250 °C and 300 °C followed by


monotonic decreases up to 12 h of hydrotreatment. It is important
to note that neither cyclopentanol nor 1-pentanol was detected in
the liquid products.
Another cyclo-ketone identified during the hydrotreatment of
bio-oil is 2-methyl cyclopentanone; its yields are shown in Fig. 4.
It is not very clear where this compound was produced from,
although the similarity of the trend observed may indicate that it
can also be a product from a compound with a structure similar
to furfural. It is important to note that cyclopentanone is also re-
ported as one of the main products of the hydrogenation of 2-
cyclopenten-1-one [15,16]. It is noteworthy that 2-cyclopenten-
1-one with and without substitution groups such as hydroxyl,
methyl/ethyl with hydroxy are commonly observed in the raw
bio-oil. Therefore it is also plausible that 2-methyl cyclopentanone
was produced from the hydrogenation of these 2-cyclopenten-1-
ones.

3.3. Carboxylic acids


Fig. 8. Yields of propyl and ethyl guaiacols as a function of hydrotreatment
temperature. Acetic acid is the most abundant (c.a. 6.75 wt%) carboxylic acid
in the raw bio-oil. Fig. 5 shows the concentration of acetic acid in
the raw bio-oil and in the hydrotreatment product for various reac-
cyclopentanol and 1-pentanol. As shown in Fig. 3a, the yield of tion temperatures and times. The concentration of acetic acid was
cyclopentanone increased with increasing reaction temperature relatively constant as expected, since much more severe hydro-
up to 250 °C, above which the yield slightly decreased. Fig. 3b treatment conditions would be required to hydrogenate carboxylic
shows that the yield of cyclopentanone reached maxima after acids [17,18].

Fig. 9. Yields of propyl and ethyl guaiacols as a function of reaction during the hydrotreatment at (a) 250 °C and (b) 300 °C.
R. Gunawan et al. / Fuel 111 (2013) 709–717 715

It is, however, interesting to note that acetic acid was not the
only carboxylic acid observed in the liquid products from the
hydrotreatment of bio-oil. Figs. 6 and 7 show that the yields of pro-
pionic, butanoic and pentanoic acids increased drastically with
increasing reaction time and temperature respectively.
To the authors’ knowledge, the evolution of C3–C5 acids during
the hydrotreatment of bio-oil was never reported in the literature
by previous investigators. The formation of formic acid, acetic acid
and levulinic acid from the hydrolysis of sugar or sugar oligomers
were reported in the literature [19]. It is hypothesised that C3–C5
carboxylic acids were possibly formed from the aliphatic chains
connecting between lignin-oligomers and/or with sugar-oligomers
in bio-oil.

3.4. Phenolics

The phenolic compounds in bio-oil are derived from the three


monolignols (the building blocks of lignin) i.e. p-coumaryl alcohol,
coniferyl alcohol and sinapyl alcohol [20]. Depending on the type
Fig. 10. Yields of methoxy phenols, phenol and methyl phenols as a function of
of biomass, these three monolignols vary in their relative abun-
hydrotreatment temperature. Yield of syringol is shown in secondary (right side) y-
axis. dance. Softwood lignin can have 90% of coniferyl alcohol, while
approximately equal proportion of coniferyl alcohol and sinapyl
alcohol are present in the hardwood lignin. Therefore, bio-oils

Fig. 11. Yields of methoxy phenols, phenol and methyl phenols as a function of reaction during the hydrotreatment at (a) 250 °C and (b) 300 °C. Yield of syringol is shown in
secondary (right side) y-axis.

Fig. 12. Effects of (a) temperature (holding time: 3 h) and (b) reaction time on the yield of toluene during the hydrotreatment at (a) 250 °C and (b) 300 °C.
716 R. Gunawan et al. / Fuel 111 (2013) 709–717

Table 3 derived from hardwood biomass have both guaiacol and syringol
Band assignment [10,11] for the FT-IR analyses of raw bio-oil and liquid products. proportionally, while the ones derived from softwood biomass
Wavenumber Functional groups are rich in syringol.
(cm 1) Figs. 8 and 9 show the yields of propyl and ethyl guaiacols as a
1767 Ester, cyclic esters lactones function of reaction temperature and time. The concentrations of
1740 Unconjugated alkyl aldehydes and alkyl esters propyl and ethyl guaiacols increased significantly with increasing
1713 Aliphatic (and fatty) acids reaction temperature, confirming their formation from the break-
1696 Unsaturated aldehydes and unsaturated ketones (very
wide band)
down of lignin-oligomer, as shown in Fig. 8. For the effects of
1654 Hydroxy unsaturated aldehydes/ketones reaction time at 250 °C, propyl and ethyl guaiacols increased sig-
1606 nificantly during the first hour followed by constant yields up to
1565 Aromatics C@C ring with various substitution 12 h of hydrotreatment, as shown in Fig. 9a. At higher tempera-
1517
ture (300 °C), similar trends were observed, but the yield de-
1501
creased substantially for the longer reaction time (see Fig. 9b).
The yields of simpler methoxy phenols (syringol, methyl guaia-
col and guaiacol), phenol and methyl phenol as a function of tem-
perature are shown in Fig. 10. The yield of all these compounds
decreased at 150 °C, but increased at higher temperatures. Similar
to propyl and ethyl guaiacols, the increasing yields of methyl guai-
acol, guaiacol and syringol must be due to the depolymerisation of
lignin-oligomer, while the decreasing yields at low temperature
might have been caused by the hydrogenation reactions. Interest-
ingly, at 300 °C, the yield of syringol decreased significantly sug-
gesting that it was hydrogenated at high temperature. Syringol
and guaiacols contain two different oxygenated functions (pheno-
lic and methoxy groups), and their hydrogenation products are cat-
echol, o-cresol, phenol, cyclohexanol, benzene, methoxybenzene,
and toluene [2,21]. In fact, Fig. 10 also shows that the yield of total
methyl phenol increased drastically at 300 °C, as well as the yield
of toluene (Fig. 12a). Similar trends can also be observed for the ef-
Fig. 13. Band areas from the spectral deconvolution of the FT-IR spectra of the raw fects of reaction time where, at 300 °C, the concentration of syrin-
bio-oil and liquid products as a function of reaction temperatures. gol decreased with increasing reaction time. In the same time, the
yield of total methyl phenol (Fig. 11b) and the yield of toluene
(Fig. 12b) also increased evidently.

3.5. FT-IR analysis

GC–MS can only identify and quantify the light molecules. It


does not give any information about high molecular mass fractions
of the bio-oil and its upgraded products. An FT-IR method has been
developed by our group [10,11] to analyse the O-containing func-
tional groups, especially carbonyl groups in the entire bio-oil
regardless of its molar mass distribution. The FT-IR spectra were
recorded for raw bio-oil and for each liquid product (aqueous
phase and oil phase, separately). The FT-IR spectra were then
deconvoluted using the nine Gaussian bands listed in Table 3
[10,11]. Figs. 13 and 14 show the results from the deconvolution
of the FT-IR spectra of raw bio-oil and liquid products (sum of
the aqueous and oil phase) for various hydrotreatment conditions.
As more water was produced during hydrotreatment (see
Fig. 1), the increases in band 1654 cm 1 could also be due to the
water bound to the molecules, as liquid water shows a medium
strength band at 1654 cm 1 in FTIR spectra due to water scissoring
frequency [22,23]. Therefore, band 1654 cm 1 was corrected for
water, and consequently all the other bands were normalised to
a water free basis, i.e. area ‘‘per g of ‘‘dry’’ bio-oil’’.
As depicted in Fig. 13, it is clear that lactones (band 1767 cm 1)
decreased during hydrotreatment, independently of the reaction
temperature. Except at 150 °C, all alkyl esters and aldehydes (band
1740 cm 1) have decreased after hydrotreatment. The increase in
these compounds at 150 °C hydrotreatment can be explained by
the conversion of unsaturated aldehydes to form saturated alde-
hydes. Band 1713 cm 1 is assigned to aliphatic acids, and
confirmed the GC–MS results that carboxylic acids were formed
Fig. 14. Band areas from the spectral deconvolution of the FT-IR spectra of the raw
during hydrotreatment (see Fig. 6). The evolution of band
bio-oil and liquid products as a function of reaction time during the hydrotreatment 1696 cm 1 (Fig. 13) can be interpreted as follows: unsaturated
at (a) 250 °C and (b) 300 °C. aldehydes and ketones were converted at 150 °C to saturated
R. Gunawan et al. / Fuel 111 (2013) 709–717 717

aldehydes/ketones. With increasing hydrotreatment temperature, References


aldehydes/ketones were converted into alcohols and even further
to alkanes and water, as the concentration of unsaturated alde- [1] Xiong WM, Fu Y, Zeng FX, Guo QX. An in situ reduction approach for bio-oil
hydroprocessing. Fuel Process Technol 2011;92:1599–605.
hydes and ketones decreased with increasing reaction tempera- [2] Elliott DC, Hart TR. Catalytic hydroprocessing of chemical models for bio-oil.
ture. The increase in band 1654 cm 1 with increasing Energy Fuels 2009;23:631–7.
hydrotreatment temperature could be ascribed to the breakdown [3] Elliott DC, Hart TR, Neuenschwander GG, Rotness LJ, Zacher AH. Catalytic
hydroprocessing of biomass fast pyrolysis bio-oil to produce hydrocarbon
of lignin oligomer at the lignin interconnecting ether bonds, with products. Environ Progr Sustain Energy 2009;28:441–9.
the formation of OH unsaturated aldehydes and ketones. As was [4] Gagnon J, Kaliaguine S. Catalytic hydrotreatment of vacuum pyrolysis oils from
also indicated by the GC–MS results, propyl guaiacol, ethyl guaia- wood. Indus Eng Chem Res 1988;27:1783–8.
[5] Wildschut J, Mahfud FH, Venderbosch RH, Heeres HJ. Hydrotreatment of fast
col, methyl phenol and phenol increased upon hydrotreatment,
pyrolysis oil using heterogeneous noble-metal catalysts. Indust Eng Chem Res
indicating the formation of these compounds as a result of the 2009;48:10324–34.
breakdown of lignin oligomer. [6] Busetto L, Fabbri D, Mazzoni R, Salmi M, Torri C, Zanotti V. Application of the
From Fig. 14, we can see that conversions of lactones (band Shvo catalyst in homogeneous hydrogenation of bio-oil obtained from
pyrolysis of white poplar: new mild upgrading conditions. Fuel
1767 cm 1) and esters and aldehydes (band 1740 cm 1) were inde- 2011;90:1197–207.
pendent of the reaction time and temperature. Acids (band [7] Garcia-Perez M, Wang S, Shen J, Rhodes M, Lee WJ, Li CZ. Effects of temperature
1713 cm 1) were formed and the yields increased with increasing on the formation of lignin-derived oligomers during the fast pyrolysis of
mallee woody biomass. Energy Fuels 2008;22:2022–32.
reaction time. The evolution of bands 1696 cm 1 and bands [8] Li X, Gunawan R, Lievens C, Wang Y, Mourant D, Wang S, et al. Simultaneous
1654 cm 1 also suggests the conversion of unsaturated alde- catalytic esterification of carboxylic acids and acetalisation of aldehydes in a
hydes/ketones and the formation of hydroxy unsaturated alde- fast pyrolysis bio-oil from mallee biomass. Fuel 2011;90:2530–7.
[9] Gunawan R, Li X, Larcher A, Hu X, Mourant D, Chaiwat W, et al. Hydrolysis and
hydes/ketones. About 50% of the unsaturated aldehydes/ketones glycosidation of sugars during the esterification of fast pyrolysis bio-oil. Fuel
were converted at 250 °C, and the conversion increased to 75% at 2012;95:146–51.
300 °C. Possibly, as stated before, at 300 °C more lignin oligomer [10] Lievens C, Mourant D, He M, Gunawan R, Li X, Li CZ. An FT-IR spectroscopic
study of carbonyl functionalities in bio-oils. Fuel 2011;90:3417–23.
was broken down with the formation of OH unsaturated alde- [11] Lievens C, Mourant D, He M, Gunawan R, Li X, Li CZ. FT-IR carbonyl bands of
hydes/ketones as a result of the breakdown of the lignin monomers bio-oils: importance of water. Fuel 2012. doi:10.1016/j.fuel.2012.01.047.
interconnecting ether bonds. [12] Venderbosch RH, Ardiyanti AR, Wildschut J, Oasmaa A, Heeres HJ. Stabilization
of biomass-derived pyrolysis oils. J Chem Technol Biotechnol 2010;85:674–86.
[13] Wildschut J, Iqbal M, Mahfud FH, Cabrera IM, Venderbosch RH, Heeres HJ.
4. Conclusions Insights in the hydrotreatment of fast pyrolysis oil using a ruthenium on
carbon catalyst. Energy Environ Sci 2010;3:962–70.
[14] Li X, Gunawan R, Wang Y, Chaiwat W, Hu X, Gholizadeh M, Mourant D, Bromly
The hydrotreatment of raw bio-oil using noble metal catalyst J, Li CZ, Upgrading of bio-oil into advanced biofuels and chemicals. Part II.
Pd/C has been studied in a batch reactor over the temperature Changes in aromatic structure and coke forming propensity during the
range of 150–300 °C. The results from the GC–MS analysis of the catalytic hydrotreatment of fast pyrolysis bio-oil with Pd/C catalyst. Fuel,
submitted for publication.
liquid products show that aldehydes were easily hydrogenated. [15] Hirota K, Touroude R, Miyasaki M, Matsumara C. A study on the hydrogenation
In particular, cyclopentanone was produced from the hydrogena- of 2-cyclopenten-1-one on nickel catalyst with deuterium by the use of
tion of furfural. Formation of propionic, butanoic and pentanoic microwave spectroscopy. J Res Inst Catal Hokkaido Univ 1977;25:37–40.
[16] Dell’Anna MM, Gagliardi M, Mastrorilli P, Suranna GP, Nobile CF.
acids were also observed in the hydrotreatment of raw bio-oil.
Hydrogenation reactions catalysed by a supported palladium complex. J Mol
The depolymerisation of the lignin-oligomer during hydrotreat- Catal A: Chem 2000;158:515–20.
ment is confirmed with the evolution of propyl guaiacol, ethyl [17] Elliott DC. Historical developments in hydroprocessing bio-oils. Energy Fuels
2007;21:1792–815.
guaiacol, methyl guaiacol, guaiacol and syringol. The hydrogena-
[18] Oasmaa A, Kuoppala E, Ardiyanti A, Venderbosch RH, Heeres HJ.
tion of the phenolic compounds also happened as shown by forma- Characterization of hydrotreated fast pyrolysis liquids. Energy Fuels
tion of phenol, methyl phenol, and toluene in the products. 2010;24:5264–72.
Deconvolution of FT-IR spectra of raw bio-oil and liquid products [19] Li N, Tompsett GA, Zhang T, Shi J, Wyman CE, Huber GW. Renewable gasoline
from aqueous phase hydrodeoxygenation of aqueous sugar solutions prepared
confirms the trends found by GC–MS. by hydrolysis of maple wood. Green Chem 2011;13:91–101.
[20] Zakzeski J, Bruijnincx PCA, Jongerius AL, Weckhuysen BM. The catalytic
Acknowledgments valorization of lignin for the production of renewable chemicals. Chem Rev
2010;110:3552–99.
[21] Nimmanwudipong T, Runnebaum RC, Block DE, Gates BC. Catalytic conversion
Australian Government funding through the Second Generation of guaiacol catalyzed by platinum supported on alumina: reaction network
Biofuels Research and Development Grant Program supports this including hydrodeoxygenation reactions. Energy Fuels 2011;25:3417–27.
[22] Abbott TP, Nabetani H, Sessa DJ, Wolf WJ, Liebman MN, Dukor RK. Effects of
project. The study also received support from the Government of bound water on FTIR spectra of glycinin. J Agric Food Chem 1996;44:2220–4.
Western Australia via the Centre for Research into Energy for Sus- [23] Mayo DW, Miller FA, Hannah RW. Course notes on the interpretation of
tainable Transport (CREST). infrared and Raman spectra. New Jersey: John Wiley & Sons; 2004.

You might also like