Sunteți pe pagina 1din 9

Bioresource Technology 133 (2013) 389397

Contents lists available at SciVerse ScienceDirect

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

Hydrothermal liquefaction of Chlorella pyrenoidosa


in sub- and supercritical ethanol with heterogeneous catalysts
Jixiang Zhang a,b, Wan-Ting Chen b, Peng Zhang b, Zhongyang Luo a, Yuanhui Zhang b,
a
b

State Key Laboratory of Clean Energy Utilization, Zhejiang University, Hangzhou 310027, China
Department of Agricultural and Biological Engineering, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA

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

" The highest mass and energy

recovery ratios of bio-crude were


71.3% and 101.8%.
" Temperature was found to be the
most dominant parameter on HTL
process.
" H2 as a processing gas improved the
bio-crude yield and quality.
" The mass ratio of the reacted ethanol
to the dry alga was about 3.26.

a r t i c l e

i n f o

Article history:
Received 20 November 2012
Received in revised form 11 January 2013
Accepted 16 January 2013
Available online 8 February 2013
Keywords:
Bio-crude
Catalyst
Ethanol
Hydrothermal liquefaction
Microalgae

a b s t r a c t
Hydrothermal liquefaction (HTL) of low lipid content microalgae Chlorella pyrenoidosa with heterogeneous catalysts was processed under sub- and supercritical conditions of ethanol (200300 C,
2.89.0 MPa, 30 min). The HTL products were separated into bio-crude, gas, solid residue and volatile
components, and then characterized. The highest mass and energy recovery ratios of bio-crude on the
dry basis of alga were 71.3% and 101.8% respectively, obtained at 240 C, while the highest higher heating
value of bio-crude was 36.19 MJ/kg, obtained at 300 C. Temperature was found to be the most dominant
parameter. H2 as a processing gas at an initial pressure of 1.03 MPa slightly improved the bio-crude yield
and quality. Raney-Ni and HZSM-5 type zeolite catalysts had no signicant effect on the presented HTL
process. The results indicated that HTL with ethanol as the solvent was able to produce 5070 wt.% of biocrude directly from C. pyrenoidosa.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
Biofuels have drawn extensive research to address the concerns
of depletion of fossil fuels, climate change and national energy
security. Algae have been addressed as a next generation biofuels
feedstock. Compared with biofuels from starch and lignocellulose,
the use of algae enjoys the advantage of higher productivity per
area per year (Demirbas and Demirbas, 2011). Additionally, algae
production does not necessarily require arable land, thus has less
Corresponding author. Address: 1304 W. Pennsylvania Avenue, Urbana, IL
61801, USA. Tel.: +1 217 333 2693; fax: +1 217 244 0323.
E-mail address: yzhang1@illinois.edu (Y. Zhang).
0960-8524/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.biortech.2013.01.076

impact on food supply. Most of the current researches are focusing


on the high lipid content microalgae for biodiesel process (Mata
et al., 2010). However, low lipid content microalgae typically have
much higher biomass yield and can grow in harsh environments
such as in wastewater, than high lipid content algae. It is worth
noting that production of low lipid content microalgae offers
opportunities to integrate wastewater treatment, carbon capture
and biofuels production into one system. Hydrothermal liquefaction (HTL) is a thermochemical process to convert biomass feedstock into an oily liquid product, known as bio-crude or bio-oil,
in and with liquid water at elevated temperature and pressure.
Compared with conventional thermochemical conversion methods
(e.g. fast pyrolysis, gasication), HTL does not require a drying pre-

390

J. Zhang et al. / Bioresource Technology 133 (2013) 389397

treatment, thus is more suitable for feedstocks with high water


content, such as algae and manure. Water plays an essential role
in HTL process. The dielectric constant (er) of water decreases from
78.85 to 13.96 when temperature increases from 25 to 350 C,
resulting in water molecular from very polar to fairly non-polar
(Archer and Wang, 1990). The dissociation constant of water (Kw)
increases from 10 14 to 10 11 just below 350 C, resulting in the
increasing rate of both acid catalyzed and base catalyzed reactions
in water (Bandura and Lvov, 2006). There have been some previous
works on producing biofuels from macroalgae (Anastasakis and
Ross, 2011; Zhou et al., 2010) and microalgae (Brown et al.,
2010; Jena et al., 2011; Minowa et al., 1995; Vardon et al., 2012;
Yang et al., 2004; Yu et al., 2011a; Yu et al., 2011b; Zou et al.,
2010) through HTL. The bio-crude yields were in the range of
2040% with higher heating values (HHV) of 2840 MJ/kg. The
oxygen and nitrogen contents of bio-crude were about 515%
and 510%, respectively.
To improve the quality of bio-crude, two routes have been considered. One approach is via catalytic hydrothermal processing
with homogeneous catalysts such as alkali catalysts or organic
acids (Anastasakis and Ross, 2011; Minowa et al., 1995; Yang
et al., 2004; Zhou et al., 2010; Zou et al., 2010), and heterogeneous
catalysts such as zeolite or supported metal catalysts (Biller et al.,
2011; Duan and Savage, 2011). The other approach is introducing
organic compounds as the processing solvents in HTL (Huang
et al., 2011; Matsui et al., 1997; Yang et al., 2011; Zhou et al.,
2012; Zou et al., 2009), among which ethanol has attracted much
of the attention. The dielectric constant of ethanol at 240 C is
4.20 (Dannhauser and Bahe, 1964). The conventional acid/base catalyzed transesterication process has been proven to undergo a
self catalyzed mechanism without the presence of catalyst in
supercritical ethanol conditions. Supercritical ethanol shows comparable physical and chemical properties to those of subcritical
water, thus replacing water with ethanol as the processing solvent
for HTL is theoretically feasible. Some noteworthy results in literatures are summarized as followed. (1) Na2CO3 was reported to increase the bio-crude yield, but not necessary for feedstock that
already contained signicant amount of sodium (Minowa et al.,
1995); (2) The gas product was mainly CO2 when processed with
alkali catalysts, while processed with organic acids produced signicant amount of CO, H2 and CH4 (Ross et al., 2010); (3) Reducing
atmosphere led to bio-crudes with a slightly increased H content
and H/C ratio (Duan and Savage, 2011); (4) Supported metal catalysts resulted in bio-crudes with a lower O/C ratio. Moreover, Pt/C
catalyst was active for the conversion of fatty acids into alkanes
(Duan and Savage, 2011); (5) A variety of fatty acid (C3C22) esters
were detected in bio-crude obtained from HTL of high lipid content
algae when processed with ethanol, indicating that ethanol acted
as a reaction substrate with algae decomposition intermediates
(Zhou et al., 2012); (6) Subcritical water and supercritical ethanol
can serve as effective hydrogen donors (Kershaw, 1997), however,
external hydrogen resource is greatly needed.

Table 1
Characteristics of C. pyrenoidosa (dry basis).
Chemical composition (wt.%)
VSa
Crude fat
Crude protein
94.4
0.1
71.3
Ash
Cellulose
Hemicellulose
5.6
0.3
0.5
Elemental composition (wt.%)
C
H
51.4
6.6
a
b

Volatile solid.
Calculated by difference.

Non-brous carbonhydrateb
22.0
Lignin
0.2
N
11.1

Ob
30.9

Chlorella pyrenoidosa is a green unicellular alga with low lipid


and high protein content that is found in both fresh and marine
waters (Becker, 1994). In this study, we combined the previous
two routes in one single process. The objective of this work was
to investigate the HTL performance by replacing water with ethanol as the solvent, and to study the effects of temperature, processing gas and catalyst on bio-crude yield and quality. HTL of C.
pyrenoidosa was carried out under sub- and supercritical
conditions of ethanol, by adjusting temperature in the range of
200300 C, under N2 or H2 atmosphere, and with/without two
catalysts, Raney-Ni and HZSM-5 type zeolite.
2. Methods
2.1. Materials
The raw material, microalgae C. pyrenoidosa, was obtained from
a health food store as food grade material (NOW Foods, Bloomingdale, IL). It was the same batch with the raw material used in our
previous works (Yu et al., 2011a,b). Characteristics of the C.
pyrenoidosa are shown in Table 1. The elemental compositions of
C. pyrenoidosa were determined using a CE 440 elemental analyzer
(Exeter Analytical, Inc., North Chelmsford, MA). The macromolecular and chemical compositions of alga were analyzed according to
the standard methods of the Association of Ofcial Analytical
Chemists (AOAC). Ammonium ZSM-5 type zeolite catalyst (Si/
Al = 50) was purchased from Zeolyst International (Conshohocken,
PA). HZSM-5 was prepared by calcination of ammonium ZSM-5 in
air at 550 C for 3 h. Aluminum-nickel alloy (Raney-nickel alloy,
50% Ni basis), supported platinum and palladium on alumina
(5 wt.% loading Pt/Al2O3, Pd/Al2O3) were obtained from SigmaAldrich. Reagent grade ethanol and acetone were obtained from
Fisher Scientic. High purity hydrogen and nitrogen were purchased from S.J. Smith Co. (Davenport, IA), and all chemicals were
used as received.
2.2. HTL process
The HTL experiments were performed in three stainless steel
cylindrical reactors of 100 ml capacity, each with a magnetic drive
stirrer (model 4593, Parr Instrument Co., Moline, IL). C. pyrenoidosa
was mixed with ethanol as the feedstock for each HTL experiment.
Typically, 20 g of the mixed feedstock with 20% of dry alga and 1 g
of catalyst (if added) were lled in the reactor. After the reactor
was sealed, high purity hydrogen (or nitrogen) gas was charged
to purge the reactor headspace ve times, and to build up to a
pre-set initial pressure of 1.03 MPa at room temperature. Then
the reactor was heated by an electric heater up to a set-point reaction temperature, and the set-point temperature was maintained
for 30 min. Stirring at 500 rpm was applied throughout the reaction. After reaction, the reactor was rapidly cooled by owing tap
water through the cooling coil located outside the reactor, and
the nal pressure was recorded at room temperature. Duplicate
or triplicate runs were performed for every experimental condition. Average values and standard deviations were reported.
2.3. Products separation and analysis
The products separation procedure is shown in Fig. 1. Once the
reactor was cooled to room temperature, the Gas products in the
reactor were carefully released through a control valve into a
Tedlar gas sampling bag (CEL Scientic CORP., Cerritos, CA). The solid/liquid products were rinsed from the reactor using twice the
amount of feedstock in acetone (40 g), and the resulted suspension
was treated in an ultrasonic bath for 30 min, followed by ltration

J. Zhang et al. / Bioresource Technology 133 (2013) 389397

391

to 800 C in 50 ml/min N2 at 10 C/min to estimate its boiling point


range distribution.
3. Results and discussions
3.1. Effects of reaction conditions on HTL products distribution
The effects of reaction conditions including temperature, processing gas and catalyst on the products yields can be explained
by Fig. 2ac, which present the results obtained at 200300 C
(from subcritical ethanol to supercritical ethanol).

Fig. 1. Products separation procedure of HTL process.

under vacuum (about 0.009 MPa) though a Whatman No. 4 lter


paper to recover the solid products. The recovered solid products
excluding the used catalyst were denoted as the Solid residue.
The solid residue was consisted of un-reacted alga, coke/char and
ash. The ltrate with the amount of 20 ml was dried under atmosphere at 75 C for 12 h to remove acetone, un-reacted ethanol
and the Volatile components, and the resulted bitumen-like
products were denoted as the Bio-crude. Yields of gas, solid residue and bio-crude, all expressed in wt.%, were calculated on the
dry basis of alga. The energy recovery of bio-crude was dened
as the energy recovered in bio-crude, and the energy recovery ratio
was dened as the ratio of the energy recovered in bio-crude to the
energy input of dry alga in the feedstock.

2.4. Characterization
The gas products were analyzed using GC (CP-3800, Varian, Palo
Alto, CA) equipped with Hayesep D 100/120 column (20 ft.  1/8
in., Alltech, Selmsdorf, Deutschland) and thermal conductivity
detector (TCD). The carrier gas was helium at a ow rate of
30 ml/min. The temperature of both the injector and detector were
120 C. The chemical compositions of the ltrate were analyzed
using GCMS (7890A, Agilent Technologies, Santa Clara, CA) with
ZB-WAX column (30 m  0.32 mm  0.25 lm, Zebron, Newport
Beach, CA) and ame ionization detector (FID). The temperature
of both the injector and detector were 250 C. The injector was in
split mode and the carrier gas was helium. The oven temperature
was controlled as such: initial temperature was 65 C for 5 min,
then heated at 5.0 C/min to 315 C and held for 5 min. Compounds
were identied according to the NIST library. The elemental compositions of the bio-crude and solid residue were determined using
a CE 440 elemental analyzer (Exeter Analytical, Inc., North Chelmsford, MA). The composition of oxygen was calculated by difference.
The HHV of the bio-crude were measured using an oxygen bomb
calorimeter (model 6200, Parr Instrument Co., Moline, IL). The
HHV of the dry alga and the solid reside were calculated using
the Dulong formula: HHV = 0.3383C + 1.442(H O/8) (Yu et al.,
2011a). Fourier transform infrared spectroscopy (FT-IR) of biocrude was performed on a FT-IR spectrometer (Thermo Nicolet
Nexus 670, Thermo Scientic, Waltham, MA) to determine its functional groups. Thermogravimetric analysis (TGA) of bio-crude was
performed on a Q50 TGA (TA Instruments, Schaumburg, IL) from 75

3.1.1. Temperature
Proteins, carbohydrates, lignocellulose and lipids are known to
undergo decomposition/depolymerization, isomerization, condensation and other reactions to form bio-crude at elevated temperature. For non-catalytic HTL under N2 environment, C. pyrenoidosa
was not fully reacted at subcritical ethanol conditions (200 and
220 C), as the solid residues yields were very high (59.5% and
37.5%, respectively). The solid residue yields decreased and
reached a minimum of 7.1% at 260 C; and then rose up slightly
to 12.0% at 300 C (Fig. 2b) due to the formation of coke/char at
higher temperature. The bio-crude yields showed the opposite
trend, reaching a maximum of 70.8% at 240 C and then drop down
to 47.1% at 300 C (Fig. 2a). Higher temperature promoted decomposition/cracking to produce small molecular fractions, part of
which were separated as gases or volatile components, and thus
resulted in the reduction of bio-crude yields. Though based on different feedstock and products separation procedure, 70.8% of biocrude yield was much higher than 2040% of bio-crude yields obtained from HTL processed with water and comparable to 72.0% of
bio-crude yield by Yang et al. (2011). The gas yields increased from
less than 1.0% to 6.1% as the temperature increased from 200 to
300 C (Fig. 2c). Similar results of increasing gas yield with increase
in temperature were reported (Anastasakis and Ross, 2011; Jena
et al., 2011; Ross et al., 2010; Yu et al., 2011a). However, the gas
yield was relatively lower compared to other reports, in which
water was used as the solvent (Vardon et al., 2012; Yu et al.,
2011a).
Yu et al. (2011b) reported that for C. pyrenoidosa the minimum
reaction temperature for bio-crude formation in water should be in
the range of 180240 C. According to the experimental data,
supercritical ethanol (>240 C) was essential to achieve the total
conversion of C. pyrenoidosa. Thus the following HTL experiments
were all conducted at the range of 240300 C.
3.1.2. Processing gas
Inter (N2 or He) atmosphere was commonly used in studies
without catalysts or with alkali catalysts. It was suggested that
both water and ethanol served as hydrogen donors under HTL conditions (Cheng et al., 2009; Kershaw, 1997). However, using high
pressure H2 in catalytic HTL with metal catalysts has attracted
much more attention (Biller et al., 2011; Duan and Savage, 2011).
As shown in Fig. 2, processing in H2 atmosphere had positive effects on HTL process, slightly increased the bio-crude yields by less
than 4% (Fig. 2a), and slightly decreased the solid residue yields at
280 and 300 C (Fig. 2b). Duan and Savage (2011) also reported
that bio-crude yields increased with the presence of high pressure
H2. The gas yields were increased when using H2 as the processing
gas (Fig. 2c).
The effects of H2 on HTL products distribution could be described as the increased amount of active hydrogen in hydrogen
donor solvent enhanced the stabilization of free radicals from the
decomposition/depolymerization of the alga, and thus prevented
retrogressive polymerization to form coke/char. The mass of

392

J. Zhang et al. / Bioresource Technology 133 (2013) 389397

3.1.3. Catalyst
Raney-Ni has high hydrogenation activity at low temperatures
and zeolite can facilitate cracking reactions that converts the large
molecular components of bio-crude into small fuel-range molecules. In this part of study, HTL with Raney-Ni and HZSM-5 zeolite
were performed under H2 and N2 atmosphere respectively. The
tested catalysts had no signicant effect on bio-crude yields in
the temperature range of 240300 C, as displayed in Fig. 2a. Furthermore, adding catalysts in HTL process tended to slightly increase the solid residue yields. Different results were reported by
Yang et al. (2011), which reported bio-crude yields increased from
34.8% to 51.6% and 72.0% with REHY and Ni/REHY catalyst respectively, under subcritical ethanol and high pressure H2 condition
(200 C). They pointed out that bifunctional catalyst Ni/REHY
achieved both acid-cracking and hydrocracking, leading to the signicant increase of bio-crude yield. Duan and Savage (2011) reported that, under subcritical water condition (350 C), bio-crude
yields over zeolite increased from 35% to 45% under helium atmosphere, but decreased from 46% to 40% under H2 atmosphere. HTL
process over Ni/SiO2Al2O3 followed the same trend. They attributed the reduced bio-crude yield to the catalytic hydrothermal gasication reactions. The insensitivity of products distribution to
catalyst in the presented work remained unclear. Rapid catalysts
deactivation was supposed to be the most inuencing factor.
It is important to note that lower bio-crude yields could be
compensated by higher quality of bio-crudes (e.g. lower oxygen
and nitrogen content, higher HHV). Further characterization of
HTL products are discussed in the following sections.
3.2. Characterization of HTL products

Fig. 2. Effects of temperature, processing gas and catalyst on products distribution


(a) bio-crude (b) solid residue (c) gas.

hydrogen in the initial pressurized H2 (0.074 g) was as much as


about 28% of that in the dry alga (0.264 g). However, only about
0.020 g of H2 was consumed during HTL process according to the
GC analysis. This might explain why the inuences of H2 atmosphere were not signicant. Compared to conventional catalytic
hydrodeoxygenation process with H2 pressure of 3.0-12.0 MPa
(Furimsky, 2000), higher initial H2 pressure was strongly
recommended.

3.2.1. HHV and energy recovery of bio-crudes


The HHV of bio-crudes in this paper were measured by calorimeter. The deviations between measured and calculated values are
within 5% as presented in the Supplementary data. The HHV and
energy recovery of bio-crudes obtained from different HTL conditions are presented in Fig. 3a and b. For non-catalytic HTL under
N2 atmosphere, the lowest HHV of bio-crudes was 28.44 MJ/kg, obtained at 220 C. The HHV increased to 34.61 MJ/kg at a HTL reaction temperature of 300 C (Fig. 3a). As a comparison, petroleum
crude has an average HHV of 42.7 MJ/kg. The HHV of bio-crudes
obtained from other HTL conditions showed the same trend. Addition of H2 slightly increased the HHV by less than 2.0 MJ/kg, since
the mass ratio of the consumed H2 to the dry alga was quite low
(0.005 g/g alga). HZSM-5 zeolite also slightly increased the HHV
of bio-crudes, while the effects of Raney-Ni were not signicant.
For non-catalytic HTL under N2 atmosphere, energy recovery of
bio-crudes reached the highest of 86.89 kJ at 240 C, then dropped
down with increase of temperature to 68.15 kJ at 300 C (Fig. 3b).
Compared to 85.36 kJ of energy input of the dry alga, the highest
energy recovery ratio was 101.8%. This result indicated that ethanol acted as a reaction substrate with alga decomposition intermediates and led to the excess energy of bio-crude. The effects of
processing gas and catalyst on energy recovery were similar with
those on bio-crude yield. H2 atmosphere was in favor of higher energy recovery. However, the elevation in energy recovery was limited by the low H2 consumption ratio. Catalysts had no signicant
effect on energy recovery of bio-crudes.
3.2.2. Elemental analysis of bio-crudes
Products from ve HTL conditions (240 C N2, 300 C N2, 300 C
H2, 300 C H2 Raney-Ni and 300 C N2 HZSM-5) were selected for
the following characterization studies. Table 2 lists the elemental
analysis of bio-crudes. Higher temperature facilitated deoxygenation as oxygen content reduced from 20.5% at 240 C to 10.9% at
300 C, and hence produced bio-crudes with higher HHV. Introduc-

J. Zhang et al. / Bioresource Technology 133 (2013) 389397

393

nents (>1% peak area) of ltrate are categorized into groups, such
as ketones (including piperidine, pyrimidinone, etc.), esters, hydrocarbons, nitrogenous compounds (including amides, amines, pyrroles, etc.) and fatty acids. The total peak area of the selected
major components accounts for >80% of the total ion chromatogram. Noteworthy, some components contain more than one functional groups are categorized in only one group, e.g., 2-pentanone,
4-amino-4-methyl- with both C@O and NH2 is categorized in ketones. The GCMS spectra and the complete chemical compositions of the ltrates are available in the Supplementary data.
According to GCMS results and possible reaction pathways
described in literatures (Peterson et al., 2008), a potential HTL
mechanism for the presented process is proposed in Fig. 4. Proteins, carbohydrates, lignocellulose and lipids rst broke down to
its corresponding monomers such as amino acids, glucose, xylose,
phenols and fatty acids under hydrothermal conditions. These
monomers further decomposed to form various types of intermediates. The monomers and their intermediates underwent a series of
reactions during the process to form the gas, liquid and solid products. As illustrated, amino acids underwent decarboxylation and
deamination reactions to form corresponding amines and keto
acids respectively. Then keto acids went through decarboxylation
to form ketones, which were abundant in the liquid products. Fatty
acids underwent decarboxylation to form aliphatic hydrocarbons,
or reacted with ethanol through esterication to form fatty acid esters. The competitive reaction of carboxyl group and ethanol to
form esters, which were more stable in hydrothermal conditions,
suppressed decarboxylation, which led to the formation of CO2,
thus decreased the gas yields. Fatty acids also reacted with amines
or ammonia through acylation to form amides. Monosaccharides
reacted with amino acids through Maillard reaction to form melanoidin (nitrogenous polymers) and solid products. Monosaccharides also further decomposed to form small molecular acids and
furfural derivatives. The acids reacted with ethanol to form esters,
while the furfural derivatives and phenols underwent repolymerization to form large molecular components and solid products.
Fig. 3. (a) Higher heating value and (b) energy recovery in bio-crudes under
different HTL conditions.

Table 2
Elemental analysis of bio-crudes.

HTL condition

Oa

240 C
300 C
300 C
300 C
300 C

61.8
71.3
71.5
70.8
70.0

8.2
8.7
9.1
8.6
8.7

9.5
9.1
8.7
9.3
8.9

20.5
10.9
10.7
11.3
12.4

N2
N2
H2
H2 Raney-Ni
N2 HZSM-5

Calculated by difference.

ing H2 atmosphere further improved the bio-crude quality, resulted in a product with an H/C ratio of 1.52 and an O/C ratio of
0.11. The nitrogen content was only slightly reduced to about
9.0% compared to that of 11.1% in the dry alga, indicating that hydrodenitrogenation was a concern for the presented HTL process.
Catalyst was found to have no signicant effect on the elemental
composition of bio-crude.

3.2.3. GCMS analysis of ltrates


The ltrate contained all liquid products from HTL reaction. GC
MS analysis of ltrate provides general information about the
chemical compositions and possible reaction pathways. The ltrate
was a complex mixture composed of various oxygenated and
nitrogenous compounds. As presents in Table 3, the major compo-

3.2.4. FT-IR analysis of bio-crudes


The FT-IR results showed good agreement with the GCMS results. The prominent C-H stretching (28502960 cm 1), CH2 and
CH3 bending (1377 and 1455 cm 1) were consistent with the significant amount of the hydrogen in bio-crude being aliphatic. The high
intensity absorption peaks at 17001730 cm 1 and 1200 cm 1 were
related to C@O stretching and CO stretching from ketones and esters. This agreed with the high peak area of ketones and esters in the
GCMS spectrum. C@O stretching (1670 cm 1) and NH bending
(1515 and 1540 cm 1) of the amide group, along with NH stretching of the amine group (near 3300 cm 1), were presented in bio-oil
obtained from 240 C, but weakened or even disappeared in bio-oils
obtained from 300 C, indicating that the fractions from proteins
decomposition underwent further decarboxylation/deamination
at higher temperature. The CO stretching at 1040 and 1100 cm 1
displayed the same trend, which could be explained as the deoxygenation went deeper at higher temperature. In addition, the
adsorption peak at 700900 cm 1 could be attributed to CH bending from aromatics and their derivatives. The FT-IR spectra are presented in the Supplementary data.
3.2.5. TG analysis of bio-crudes
TGA applied in simulated distillation is regarded as a miniature
distillation. Although some thermal degradation is likely, TGA
provides an estimate of the boiling range of bio-crude (Ross
et al., 2010). TG/DTA curves of bio-crudes are presented in Fig. 5.
Heating of bio-crudes under an inert atmosphere to 800 C resulted
in a weight loss of about 85%. The weight loss of bio-crudes before
150 C was about 2%, indicating that the drying process effectively

394

J. Zhang et al. / Bioresource Technology 133 (2013) 389397

Table 3
Major components of ltrates.
Compound name

Ketones
3-Penten-2-one, 4-methyl
Cyclopentanone, 2-methyl
2-Pentanone, 4-hydroxy-4-methyl
2-Pentanone, 4-amino-4-methyl
2-Cyclopenten-1-one, 2,3-dimethyl
2,2,6,6-Tetramethyl-4-piperidone
1,3-Dimethyl-3,4,5,6-tetrahydro-2(1H)-pyrimidinone
Pyrrolo[1,2-a]pyrazine-1,4-dione,hexahydroEsters
Ethyl acetate
Propanoic acid, ethyl ester
Pentanedioic acid, diethyl ester
Benzenepropanoic acid, ethyl ester
Hexadecanoic acid, ethyl ester
Ethyl 9-hexadecenoate
Octadecanoic acid, ethyl ester
Ethyl oleate
Linoleic acid ethyl ester
9,12,15-Octadecatrienoic acid, ethyl ester, (Z,Z,Z)
DL-Proline, 5-oxo, methyl ester
9,12,15-Octadecatrienoic acid, methyl ester, (Z,Z,Z)
Hydrocarbons
Cyclopentene,1-hexyl
Cyclooctene, 1,2-dimethyl
2-Hexadecene, 3,7,11,15-tetramethyl-, [R-[R,R-(E)]]
5-Undecene, 3-methyl, (E)
Nitrogenous compounds
Hexadecanenitrile
1-Hexanamine
Hexadecanamide
1H-Pyrrole, 3-ethyl-2,4,5-trimethyl
D-Glucitol, 1-deoxy-1-(heptylamino)
Fatty acids
9,12,15-Octadecatrienoic acid, (Z,Z,Z)
Total
a

Peak area (%)a


240 C N2

300 C N2

300 C H2

300 C H2 Raney-Ni

300 C N2 HZSM-5

36.84
4.37

58.67
9.59

13.47

31.36

1.03
15.28
1.48
1.21
33.62
1.75
0.84
0.43
0.32
7.80
1.40
1.94
1.51
6.48
4.97
5.13

2.77
14.55
0.40

45.91
6.58
1.49
21.44
1.37
1.28
13.24
0.51

45.54
3.74
7.61
10.42
5.92
1.44
15.68
0.73

45.20
5.79
3.44
20.29
1.77
1.41
11.84
0.66

16.01
1.33
0.78
0.87
1.10
5.06
0.93
1.12
2.01
1.90
0.25
0.66

22.55
1.86
1.02
1.17
1.17
6.56
1.09
3.18
2.90
2.71
0.19
0.70

25.92
1.85
1.16
1.19
1.37
7.99
1.44
1.77
3.36
3.31
0.94
1.54

25.20
2.17
1.15
1.13
1.39
7.30
1.34
1.77
2.92
4.05
0.80
1.18

7.60
5.51
0.74
0.93
0.42
1.00
0.68
0.15
0.17

3.80
1.71
0.92
1.17

5.20
2.66
1.25
1.29

8.19
0.53
0.17
0.43
7.06

7.36
4.55
1.24
1.01
0.56
2.66
1.24
0.33
0.33
0.76

83.28

80.45

81.48

80.83

1.05
6.58
1.97
2.51
0.32
1.78
5.72
1.05
1.75

5.23
0.45
0.21
0.24
4.33

2.92
3.16
3.16
85.92

Excluding ethanol and acetone.

Fig. 4. Potential reaction mechanism of the presented HTL process. (1) hydrolysis; (2) decarboxylation; (3) deamination; (4) esterication; (5) polyerization; (6)
decomposition; (7) Maillard reaction; (8) acylation.

removed the volatile components. Bio-crude obtained from lower


temperature had a maximum weight loss rate at around 300 C,
while bio-crudes obtained from higher temperature showed two
weight loss peaks at around 220 and 420 C, respectively (Fig. 5).

The TG/DTA curves of bio-crudes obtained from 300 C matched


well with each other, indicating that temperature was the most
critical parameter. Table 4 lists the boiling point distribution of
bio-crudes. Bio-crudes obtained from 300 C had less distribution

395

J. Zhang et al. / Bioresource Technology 133 (2013) 389397

Fig. 5. TG/DTA curves of bio-crudes.

in 250370 C, but more distribution in <250 and >370 C, compared to bio-crude obtained from 240 C. It was consistent with
the previous discuss that, at higher temperature, bio-crude underwent decomposition and repolymerization to form smaller and larger molecular components, respectively. All bio-crudes contained
signicant amount of high boiling fractions.
3.3. Ethanol consumption and mass/energy balance
Ethanol acted as both a solvent and a reaction substrate with
the alga decomposition intermediates, as discussed previously.
However, the consumption of ethanol was rarely reported in literatures. Zhang et al. (2012) and Dang et al. (2013) examined ethanol
consumption and its effects on the mass and energy balance, in
which supercritical ethanol was used for the upgrading of pyrolysis
bio-oil. When the mass ratio of ethanol to pyrolysis bio-oil in the
feedstock was 5.0, noting that the pyrolysis bio-oil contained signicant amount of water (41.8%), the mass ratio of the reacted ethanol to the pyrolysis bio-oil (excluding water) was about 1.9 (Dang
et al., 2013).
In this study, the ethanol content of the ltrates was determined by GCMS using the ethanol calibration curve, and the ethanol consumption was obtained by subtraction of un-reacted
ethanol from initial ethanol input. The calculated mass/energy balances from ve HTL conditions are listed in Table 5. The H2 was not
taken into consideration since the amount of the consumed H2 was
negligible. The mass ratio of the reacted ethanol to the dry alga at
240 and 300 C were 3.17 and 3.28 respectively. More ethanol was
consumed and more volatile components were produced at higher
HTL reaction temperature. Furthermore, the increment of the

volatile components formation (1.3 g) was higher than that of the


ethanol consumption (0.46 g), indicating that bio-crude underwent
decomposition/cracking to form small molecular fractions, which
agreed with the conclusions drawn from the characterization studies. The energy balance was consistent with the products distribution and the HHV data. Gaseous products had almost no heating
value since it contained over 98 wt.% of CO2. Solid residues had
HHV from 25 to 30 MJ/kg, and accounted for less than 3 kJ of the
energy output in one batch of experiment. The energy recovered
in bio-crude was higher than the energy input of dry alga at
240 C. The energy output of bio-crude and volatile components
together accounted for over 98% of the total energy output.
The high ethanol consumption ratio was a concern for the presented HTL process, as ethanol is a good fuel/fuel additive itself.
The authors proposed that the ethanolwater solution with optimized ethanol/water ratio might serve as an appropriate solvent
for the presented HTL process and solve this high ethanol consumption issue. Furthermore, the large amount of volatile components is in need of recycle and characterization in future works.
3.4. Upgrading of bio-crudes and volatile components
In this part of study, a two-step process, rst a HTL process at
lower temperate then a catalytic upgrading process at higher temperature, was investigated and compared with direct HTL process.
HTL of C. pyrenoidosa (40 g of feedstock amount with 40% of dry
alga load) with ethanol as the solvent was carried out under the
condition of 260 C, 60 min in H2. After the reactor was opened,
the solid/liquid products were rinsed from the reactor using 40 g
of ethanol rather than acetone. The ltrate was obtained by follow-

Table 4
Boiling point distribution of bio-crudes.
Distillate range (C)

<200
200250
250300
300370
>370

Boilling point range of bio-crudes (wt.%)


240 C N2

300 C N2

300 C H2

300 C H2 Raney-Ni

300 C N2 HZSM-5

11.7
12.8
20.1
25.0
30.4

11.6
19.8
14.3
14.6
39.7

10.7
20.5
15.0
15.3
38.5

10.4
19.3
14.3
15.6
40.4

10.9
19.8
15.2
16.0
38.1

396

J. Zhang et al. / Bioresource Technology 133 (2013) 389397

Table 5
Mass and energy balances of ve HTL conditions.
HTL condition

Input (g)

Output (g)

Dry alga

Moisture

240 N2
300 N2
300 H2
300 H2 Raney-Ni
300 N2 HZSM-5
HTL condition

4.00
4.00
4.00
4.00
4.00
Input (kJ)

0.28
0.28
0.28
0.28
0.28

240 N2
300 N2

Dry alga
85.36
85.36

Moisture
0.00
0.00

Ethanol
15.72
15.72
15.72
15.72
15.72
Ethanol
466.88
466.88

Bio-crude

Solid residue

Gas

2.90
1.97
2.02
1.96
2.07
>Output (kJ)

0.69
0.45
0.42
0.57
0.49

0.09
0.23
0.53
0.52
0.15

Bio-crude
86.89
68.15

Solid residue
2.19
1.74

Gas
0.00
0.00

Volatile componentsa
13.27
14.79
14.22
14.60
14.65
Volatile componentsa
459.51
474.39

Unreacted ethanol
3.05
2.56
2.81
2.35
2.64
Unreacted ethanol
90.54
76.11

Caclulated by difference.

ing the same separation procedure. Then, 20 ml of the ltrate was


dried to determine the bio-crude yield and for further analysis.
Meanwhile, 20 g of the ltrate was loaded with 1 g of Pt/Al2O3 or
Pd/Al2O3 catalysts separately into the reactor for the upgrading
experiments. Upgrading of the HTL ltrates (bio-crudes and volatile components) in supercritical ethanol with catalysts were
performed under the condition of 300 C, 60 min in H2. The
upgrading products followed the same products separation procedure as described in Section 2.3, and the upgraded bio-crude was
obtained and analyzed.
The rst-step HTL process had a bio-crude yield of 65.5 0.7%,
and the bio-crude had an HHV of 32.43 1.07 MJ/kg. The upgrading process resulted in an overall upgraded bio-crude yield of
49.8 0.6%, and the upgraded bio-crude had an HHV of
36.40 0.40 MJ/kg. Compared to direct HTL of C. pyrenoidosa under
300 C, 30 min in H2 (50.7 0.2% and 36.19 0.16 MJ/kg), the
upgrading process was found to have no signicant effect on products distribution or HHV of bio-crude. However, the upgraded biocrude had less distribution in >250 C (60.8% and 60.5% compared
to 68.8%), and more distribution in <200 C (19.1% and 20.6% compared to 10.7%), according to the TG/DTA analysis presented in the
Supplementary data.

4. Conclusions
The results have demonstrated the feasibility of directly producing bio-crude from low lipid content microalgae through HTL
process with ethanol as the solvent. Supercritical ethanol condition
(>240 C) is essential for the conversion of C. pyrenoidosa, and higher temperature facilitates deoxygenation. H2 as the processing gas
slightly improves the bio-crude yield and quality, while catalysts
have no signicant effect. The characterization results are consistent, based on which a potential HTL mechanism is proposed.
Upgrading process improves the boiling point distribution of biocrude. However, high ethanol consumption and denitrogenation
remain concerns for the presented HTL process.

Acknowledgements
The authors appreciate the nancial support from China Scholarship Council to sponsor the living expense for study abroad; and
the University of Illinois for providing the experiment facilities and
supplies for the research. Acknowledgment also goes to fellow
graduate students and researchers in the Bioenvironmental Engineering Division at University of Illinois at Urbana-Champaign,
for their assistance during the project.

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.
01.076.

References
Anastasakis, K., Ross, A.B., 2011. Hydrothermal liquefaction of the brown macroalga Laminaria Saccharina: effect of reaction conditions on product distribution
and composition. Bioresource Technology 102 (7), 48764883.
Archer, D.G., Wang, P.M., 1990. The dielectric-constant of water and Debye-Huckel
limiting law slopes. Journal of Physical and Chemical Reference Data 19 (2),
371411.
Bandura, A.V., Lvov, S.N., 2006. The ionization constant of water over wide ranges of
temperature and density. Journal of Physical and Chemical Reference Data 35
(1), 1530.
Becker, E.W., 1994. Microalgae: Biotechnology and Microbiology. Cambridge
University Press, Cambridge.
Biller, P., Riley, R., Ross, A.B., 2011. Catalytic hydrothermal processing of microalgae:
decomposition and upgrading of lipids. Bioresource Technology 102 (7), 4841
4848.
Brown, T.M., Duan, P.G., Savage, P.E., 2010. Hydrothermal liquefaction and
gasication of Nannochloropsis sp. Energy & Fuels 24, 36393646.
Cheng, Z.M., Ding, Y., Zhao, L.Q., Yuan, P.Q., Yuan, W.K., 2009. Effects of supercritical
water in vacuum residue upgrading. Energy & Fuels 23, 31783183.
Dang, Q., Luo, Z.Y., Zhang, J.X., Wang, J., Chen, W., Yang, Y., 2013. Experimental study
on bio-oil upgrading over Pt/SO42-/ZrO2/SBA-15 catalyst in supercritical
ethanol. Fuel 103, 683692.
Dannhauser, W., Bahe, L.W., 1964. Dielectric constant of hydrogen bonded liquids 3.
Superheated alcohols. Journal of Chemical Physics 40 (10), 3058.
Demirbas, A., Demirbas, M.F., 2011. Importance of algae oil as a source of biodiesel.
Energy Conversion and Management 52 (1), 163170.
Duan, P.G., Savage, P.E., 2011. Hydrothermal liquefaction of a microalga with
heterogeneous catalysts. Industrial & Engineering Chemistry Research 50 (1),
5261.
Furimsky, E., 2000. Catalytic hydrodeoxygenation. Applied Catalysis a-General 199
(2), 147190.
Huang, H.J., Yuan, X.Z., Zeng, G.M., Wang, J.Y., Li, H., Zhou, C.F., Pei, X.K., You, Q.A.,
Chen, L.A., 2011. Thermochemical liquefaction characteristics of microalgae in
sub- and supercritical ethanol. Fuel Processing Technology 92 (1), 147153.
Jena, U., Das, K.C., Kastner, J.R., 2011. Effect of operating conditions of
thermochemical liquefaction on biocrude production from Spirulina platensis.
Bioresource Technology 102 (10), 62216229.
Kershaw, J.R., 1997. Comments on the role of the solvent in supercritical uid
extraction of coal. Fuel 76 (5), 453454.
Mata, T.M., Martins, A.A., Caetano, N.S., 2010. Microalgae for biodiesel production
and other applications: a review. Renewable & Sustainable Energy Reviews 14
(1), 217232.
Matsui, T.O., Nishihara, A., Ueda, C., Ohtsuki, M., Ikenaga, N.O., Suzuki, T., 1997.
Liquefaction of micro-algae with iron catalyst. Fuel 76 (11), 10431048.
Minowa, T., Yokoyama, S., Kishimoto, M., Okakura, T., 1995. Oil production from
algal cells of Dunaliella-Tertiolecta by direct thermochemical liquefaction. Fuel
74 (12), 17351738.
Peterson, A.A., Vogel, F., Lachance, R.P., Froling, M., Antal, M.J., Tester, J.W., 2008.
Thermochemical biofuel production in hydrothermal media: a review of suband supercritical water technologies. Energy & Environmental Science 1 (1), 32
65.
Ross, A.B., Biller, P., Kubacki, M.L., Li, H., Lea-Langton, A., Jones, J.M., 2010.
Hydrothermal processing of microalgae using alkali and organic acids. Fuel 89
(9), 22342243.

J. Zhang et al. / Bioresource Technology 133 (2013) 389397


Vardon, D.R., Sharma, B.K., Blazina, G.V., Rajagopalan, K., Strathmann, T.J., 2012.
Thermochemical conversion of raw and defatted algal biomass via
hydrothermal liquefaction and slow pyrolysis. Bioresource Technology 109,
178187.
Yang, C., Jia, L.S., Chen, C.P., Liu, G.F., Fang, W.P., 2011. Bio-oil from hydroliquefaction of Dunaliella sauna over Ni/REHY catalyst. Bioresource Technology
102 (6), 45804584.
Yang, Y.F., Feng, C.P., Inamori, Y., Maekawa, T., 2004. Analysis of energy conversion
characteristics in liquefaction of algae. Resources Conservation and Recycling
43 (1), 2133.
Yu, G., Zhang, Y., Schideman, L., Funk, T.L., Wang, Z., 2011a. Hydrothermal
liquefaction of low lipid content microalgae into bio-crude oil. Transactions of
the Asabe 54 (1), 239246.
Yu, G., Zhang, Y.H., Schideman, L., Funk, T., Wang, Z.C., 2011b. Distributions of
carbon and nitrogen in the products from hydrothermal liquefaction of lowlipid microalgae. Energy & Environmental Science 4 (11), 45874595.

397

Zhang, J.X., Luo, Z.Y., Dang, Q., Wang, J., Chen, W., 2012. Upgrading of bio-oil over
bifunctional catalysts in supercritical monoalcohols. Energy & Fuels 26 (5),
29902995.
Zhou, D., Zhang, L.A., Zhang, S.C., Fu, H.B., Chen, J.M., 2010. Hydrothermal
liquefaction of macroalgae Enteromorpha prolifera to bio-oil. Energy & Fuels
24, 40544061.
Zhou, D., Zhang, S.C., Fu, H.B., Chen, J.M., 2012. Liquefaction of macroalgae
Enteromorpha prolifera in sub-/supercritical alcohols: direct production of
ester compounds. Energy & Fuels 26 (4), 23422351.
Zou, S.P., Wu, Y.L., Yang, M.D., Kaleem, I., Chun, L., Tong, J.M., 2010. Production and
characterization of bio-oil from hydrothermal liquefaction of microalgae
Dunaliella tertiolecta cake. Energy 35 (12), 54065411.
Zou, S.P., Wu, Y.L., Yang, M.D., Li, C., Tong, J.M., 2009. Thermochemical catalytic
liquefaction of the marine microalgae Dunaliella tertiolecta and
characterization of bio-oils. Energy & Fuels 23 (7), 37533758.

S-ar putea să vă placă și