You are on page 1of 9

Bioresource Technology 219 (2016) 643651

Contents lists available at ScienceDirect

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

Promotion of hydrogen-rich gas and phenolic-rich bio-oil production


from green macroalgae Cladophora glomerata via pyrolysis over its
bio-char
Omid Norouzi a, Sajedeh Jafarian a, Farid Safari b, Ahmad Tavasoli a,b,, Behnam Nejati a
a
b

School of Chemistry, College of Science, University of Tehran, Tehran, Iran


Department of Energy Engineering, Science and Research Branch, Islamic Azad University, Tehran, Iran

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

 Conversion of Cladophora glomerata

via pyrolysis for biofuel production.


 Non-catalytic tests to determine the

optimum condition for bio-oil


production.
 FESEM, ICP and BET method for
characterization of algal bio-char.
 Upgrading of bio-oil in the presence
of algal bio-char as a catalyst.
 An increase in H2 concentration and
selectivity by the addition of algal
bio-char.

a r t i c l e

i n f o

Article history:
Received 26 June 2016
Received in revised form 3 August 2016
Accepted 5 August 2016
Available online 8 August 2016
Keywords:
Cladophora glomerata
Pyrolysis
Bio-char
Bio-oil
Macroalgae

a b s t r a c t
Conversion of Cladophora glomerata (C. glomerata) as a Caspian Seas green macroalgae into gaseous, liquid and solid products was carried out via pyrolysis at different temperatures to determine its potential
for bio-oil and hydrogen-rich gas production for further industrial utilization. Non-catalytic tests were
performed to determine the optimum condition for bio-oil production. The highest portion of bio-oil
was retrieved at 500 C. The catalytic test was performed using the bio-char derived at 500 C as a catalyst. Effect of the addition of the algal bio-char on the composition of the bio-oil and also gaseous products was investigated. Pyrolysis derived bio-char was characterized by BET, FESEM and ICP method to
show its surface area, porosity, and presence of inorganic metals on its surface, respectively. Phenols were
increased from 8.5 to 20.76 area% by the addition of bio-char. Moreover, the hydrogen concentration and
hydrogen selectivity were also enhanced by the factors of 1.37, 1.59 respectively.
2016 Elsevier Ltd. All rights reserved.

1. Introduction
Industrial development during the last centuries has made a
leap in the progress of human life. However, recently, the alarming
effects of this development on the environment is being appeared
dramatically (Dincer and Rosen, 2012; Akalin et al., 2012). Climate
change and energy security are among the most important con Corresponding author at: School of Chemistry, College of Science, University of
Tehran, Tehran, Iran.
E-mail addresses: tavasoli.a@ut.ac.ir, a-tavasoli@srbiau.ac.ir (A. Tavasoli).
http://dx.doi.org/10.1016/j.biortech.2016.08.017
0960-8524/ 2016 Elsevier Ltd. All rights reserved.

cerns of the recent era (Jones and Warner, 2016; Shoja et al.,
2013). Production of energy carriers with a minimum carbon content in the combustion process would be promising (Ali et al.,
2013). Moreover, availability and renewability of the primary
resources are crucial. In the recent decade, extensive research
has been carried out on the production of gaseous and liquid
energy carriers from bio-renewable resources (Krishnan and
McCalley, 2016). Biomass feedstocks are rich in carbon and hydrogen and can be converted into valuable fuels and chemicals
(Phillips et al., 2016). However, agricultural wastes as the second
generation of feedstocks are dependent on landing availability

644

O. Norouzi et al. / Bioresource Technology 219 (2016) 643651

and have some competition with food production (Guragain et al.,


2016). The third generation of feedstock may overcome the shortcomings associated to the first and second generation feedstocks.
Utilization of algal biomass instead of other terrestrial feedstocks
can reduce the land use and water consumption in the cycle of
bioenergy production (Savage, 2012; Singh et al., 2011). Macroalgae do not require land or fresh water for cultivation (Vassilev
and Vassileva, 2016). In the estimation of Chung et al. (2011), cultivation of macroalgae along the coastlines can annually capture
one billion tons of carbon. However, a lower amount of carbohydrates may results in lower value products compared with agricultural wastes but, the presence of lipid and protein in the
macroalgae and lower amount of lignin in their structure can make
the conversion process much easier and less energy intensive
(Safari et al., 2016a). C. glomerata has a widespread distribution
in Caspian Sea Coast, which results in environmental damages
and are a cause of a eutrophication of the water (Soltani et al.,
2014). Thus, there is a big social demand to explore an efficient
and cost effective process for conversion of this bio-resource
(Mihranyan, 2011). Holistic analysis of the Pyrolysis products of
C. glomerata as a widespread green alga has not been investigated
before. Plis et al. (2015) studied the thermochemical properties of
this strain for possible conversion to value-added product. Due to
the high content of volatile matter, pyrolysis process was strongly
recommended for producing liquid and gaseous fuels. The liquid
bio-oil from pyrolysis of algae is CO2 neutral and environmentfriendly. However, the crude bio-oil can be upgraded via a catalytic
process for the production of transportation fuel. In addition, the
syngas derived from pyrolysis can be applied for both hydrogen
and power production in the industry (Jafarian et al., 2016; Safari
et al., 2016b). And to date, various bio-chars, particularly those
derived from terrestrial biomass have been applied as a soil
amendment, water treatment, and fertilizer (de la Rosa et al.,
2014). However, studies using bio-char as a catalyst in biomass
pyrolysis and bio-oil upgrading have been lacking. Ren et al.
(2014) investigated the effects of Corn stover bio-char as a catalyst
in biomass catalytic pyrolysis and bio-oil upgrading. Recently, the
use of algal bio-char has gained extensive attention for various
applications because marine macroalgae are massively abundant
and bio-char converted from this source has a relatively high
porosity and surface area, and it contains higher ash and functional
groups on the surface than the bio-char that derived from terrestrial biomass (Jung et al., 2016). These properties make the algal
bio-char interesting as catalyst support and catalyst. No published
data corresponding to investigating the effects of algal bio-char as
a catalyst for the production of bio-oil, syngas, and bio-char through
pyrolysis of macroalgae is available. Several studies on pyrolysis of
algal biomass have been reported in the literature, including Microcystis sp. and S. platensis (Li et al., 2012). Omoriyekomwan et al.
(2016) reported microwave pyrolysis of palm kernel shell using activated carbon (AC) and lignite char (LC) as catalysts. The maximum
concentration of phenol in bio-oil was 64.58 (%Area). Yildiz et al.
(2015) investigated the effect of biomass ash on the fast pyrolysis
of pinewood. The presence of inorganic compounds on the ash promoted the conversion of phenols and suppressed the conversion of
sugars and acids. Song Hu et al. (2015) reported the effect of alkali
and alkaline earth metals (AAEMs) on the promotion of hydrogen
via an increase in the absorption of H2O molecule. These mineral
nutrients are a part of the biomass structure, bound at hydroxyl
and/or phenolic groups in the form of cations or as a salt. In this
study, the focus has been given on upgrading of bio-oil and
improvement of hydrogen-rich gas production through pyrolysis
of C. glomerata via the addition of its bio-char as a catalyst.
C. glomerata offers a promising feedstock for production bioenergy using thermochemical technology. In the previous work of
the authors, it was shown that Enteromorpha intestinalis as a Caspian

Sea macroalgae had a great potential for hydrogen production


(Norouzi et al., 2016). In this study, another major algal biomass of
southern coast of Caspian Sea was investigated to determine its
potential for hydrogen-rich gas and bio-oil production. To the best
of our knowledge, there are no previous work in the literature on
the detailed and holistic investigation of the pyrolysis of C. glomerata
macroalgae. The main objective of this study is to investigate the
effect of bio-char on the conversion of macroalgae to gaseous and
liquid products. The main novel investigation of the current study
were to:
 Holistic characterization of C. glomerata as widespread macroalgae found in the Caspian Sea coast of Iran.
 Investigate of the gaseous, liquid and solid products of
C. glomerata as a third generation bioresource for the first time
to determine it potential for further technical development.
 Characterize and utilize the pyrolysis derived solid bio-char of
C. glomerata as a catalyst for enhancement in Hydrogen production and upgrading the bio-oil.
2. Material and methods
2.1. Feedstock preparation
The selected macroalgae used in the present study was
C. glomerata, which was collected from Sisangan area located in
Southern Caspian Sea coast, Iran, where a stable coverage was
found. It was dried under atmospheric conditions for 48 h and
ground to the particle size <150 lm in diameter.
2.2. Feedstock characterization
The proximate analysis for determination of volatile matter and
moisture content of the C. glomerata was performed by thermogravimetric analysis (TGA). The experiments were carried out by
feeding the algae to a thermogravimetric analyzer (TGA/SDTA851
and METTLER-TOLEDO compact). The tested conditions were controlled under the nitrogen atmosphere at the temperatures in the
range of 30900 C with a heating rate of 10 C min1. The temperature was kept constant for 5 min before increasing it up to 900 C
at a rate of 100 C min1 for measurement of ash content the
method of Sluiter et al. (2008), was adapted to biochar samples.
Briefly, after drying at 105 C for 5 h, the samples were placed in
a porcelain crucible and heated in a muffle furnace at 575 25 C
for 24 6 h to constant weight in a muffle furnace. Each sample
was analyzed in triplicate. Fixed carbon was calculated via Eq. (1).

Fixed carbon% 100  Ash%  volatile matter%  moisture%

The elemental analysis of the biomass sample was performed using


a CHNS analyzer (Vario ELIII by Elementar, Germany). Also, the percentage of oxygen was determined through the balance via Eq. (2).

O% 100  C%  H%  N%  S%  Ash%

The amount of protein in the sample was recognized in accordance with National Standard in China GB/T 6432-1994 by the
method of Kjeldahl, while the amount of lipid was determined
via the method of solvent extraction in accordance with National
Standard in China GB/T 6433-2006 (Zhao et al., 2015). Carbohydrate content was calculated by mass balance method by Eq. (3).

Carbohydrate% 100  Protein%  lipid%  ash%

In order to determine the physical properties of the raw


C. glomerata, surface analysis and porous textural analysis of the
material was measured by an ASAP-2010 system from micromeritics. The samples were degassed at 200 C for 4 h under 50 mTorr
vacuum and their BET area and pore volume were determined.

O. Norouzi et al. / Bioresource Technology 219 (2016) 643651

2.3. Experimental setup and procedure


Fig. 1, shows the reaction setup. As shown, the reaction was
performed in a fixed bed reactor made of a quartz glass tube with
800 mm long and 10 mm internal diameter. Quartz wool, feed, and
catalyst bed were fixed in the center of the reactor. The reactor was
located inside a tubular furnace with 20 mm ID and 200 mm OD
with the total heating length of 700 mm. The furnace temperature
was controlled using an electrical heater and a PID temperature
controller and measured using a K-type thermocouple. Argon
was used as inert carrier gas with exactly 30 ml/min flow rate
using mass flow controller. The reaction products included char,
gas and bio-oil compounds. The bio-oil was collected in a liquid
trap of ice-salt bath and the gaseous fraction was collected over
a sodium chloride saturated brine solution in a graduated column
and the char remained in the center of the reactor. For the catalytic
tests, the biochar obtained from the pyrolysis at the optimum condition, were used as a catalyst. This biochar was blended as a catalyst with biomass sample and fed into the reactor. After the
catalytic process, the certain amount of the collected biochar
(equals to the non-catalytic biochar) was added back to the reactor
for the next experiment. To better understanding of the process
and clarifying the method of this study. This process is selfpromoted by giving a feedback from its own products. 1 g of
C. glomerata is dried first, and then, sieved and chopped for
pyrolysis. The process starts with non-catalytic pyrolysis which
the amount of the obtained bio-char is 0.4 g. Then, catalytic pyrolysis in the presence of this bio-char is carried out. Catalytic process
results in 0.75 g overall bio-char (0.4 g as a catalyst and 0.35 from
the process). In this case, 0.4 g of bio-char is added back to the
reactor as a catalyst each time, and the remaining, which is equal
to 0.35 g, is collected for further usage such as soil amendment.

645

For carbon balance, the amount of carbon in the bio-oil and biochar was analyzed by a TOC analyzer (Shimadzu, TOC-L) and a
CHNS analyzer, respectively. Moreover, the amount of carbon in
the gas phase was calculated as the total moles of carbon in
carbon-containing gaseous products in mass basis.
2.4. Product analysis
The volume of collected gasses was measured from the displacement of the solution in the column. The quality and quantity
of produced gaseous mixture were analyzed using gas chromatography technique (Varian 3400 and Teyfgostar-Compact). The biooil in glass condenser and char left inside the reactor were
measured by measuring the difference in weights of glass condenser and reactor before and after the reaction. Considering the
small reactor volume and feedstock quantity added into the system, reported data are the averages of several observations for
each experiment in order to make results more statistically confident. Also, the accuracy of data collection and comparison between
means of compositions and yields from the processing of two samples was studied through ANOVA (analysis of variance). The statistical level was 5% with p < 0.05.
2.4.1. Gaseous product analysis
Gas samples were taken by tight syringes and injected into Gas
chromatographs column. A gas chromatograph (Varian 3400 and
Teyfgostar-Compact) had been equipped with PORAPAK Q-S
80/100 (30 m long, 0.53 mm I.D) column, a methanizer and Flame
Ionization Detector (FID). Argon was used as carrier gas and oven
temperature program was in the following: 40 C isothermal for
5 min, an increase in the temperature from 40 to 75 C in
17.5 min and isothermal in 75 C for 5 min. The methanizer option

Fig. 1. Schematic diagram of the pyrolysis experimental setup.

646

O. Norouzi et al. / Bioresource Technology 219 (2016) 643651

enables the FID to detect levels of CO and CO2. During analysis,


methanizer is heated to 380 C with the FID detector body. When
the column effluent mixes with the FID hydrogen supply and
passes through the methanizer, CO and CO2 are converted to
methane. GC was calibrated with standard gas mixture supplied
by Roham Company in Tehran, Iran. The standard deviation for
the results of gas composition was calculated to be 2%.
2.4.2. Liquid product analysis
The chemical composition of bio-oil was analyzed using gas
chromatography/mass spectrometry (GC/MS, Agilent Technology
(HP)). An HP-5 column (30 m  0.25 mm i.d., 0.25 lm film thickness) was applied to the analysis. Helium (99.999%) was used as
the carrier gas at a constant flow rate of 1 mL min1. The oven temperature was programmed from 50 to 250 C at the heating rate of
10 C min1. MS was conducted under the following operational
conditions: transfer line 23 C, ion source 230 C, and electron
energy 70 eV.
2.4.3. Solid product analysis
Biochar derived from pyrolysis of C. glomerata, was elementally
analyzed by a CHNS analyzer. Also, a Field Emission Scanning Electro Microscope (FESEM MIRA3 LM, Tuscan), provided by Nanostructure and coating Lab, Sharif university of technology, was
used for scanning the morphology of the surface for raw feedstock
and bio-char. Coating of the sample is required in the field of electron microscopy to enable and improve the imaging of samples. All
samples in this study were coated with a thin layer of gold to create a conductive layer of metal which inhibits charging, reduce
thermal damage and improve the secondary electron signal
required for topographic characterization. The surface area, total
pore volume of the C. glomerata bio-char were determined as mentioned in Section 2.2.
2.5. Data interpretation
Concentration of different pyrolysis products has been presented in terms of volume concentration which is presented by
Eq. (4):

Concentration v=v%

Volume of the intended Product


Total volume of the pyrolysis product
4

CGE

Carbon in gaseous products


Carbon in Cladophora

Carbon balance CGE CLE CSE

9
10

3. Results and discussion


3.1. Results
3.1.1. Biomass characterization
The elemental analysis of C. glomerata was carried out to determine the weight percentage of each element. The results were as
follows: 34.4 wt% C, 5 wt% H, 5.2 wt% N, 2.3 wt% S, 18.6 wt% O.
The heating value of the C. glomerata is about 14.97 MJ/kg and is
relatively high compared to other agricultural residues (Nanda
et al., 2016). The TGA curves and DTG (Fig. 2) shows that the thermal decomposition of C. glomerata occurs in three steps. The first
stage occurs in the 30200 C which corresponds for the drying,
and slight volatile release. The second stage of decomposition,
ranging from 200 to 500 C with the major weight loss (46%), is
associated to the major pyrolysis in which devolatilization takes
place. And the last stage corresponds to solid residue. The main
three picks in DTG diagram are corresponding to the water and
water loosely bound to biomolecules (46 C), thermal cracking of
protein and carbohydrates (270 C) and lipids non-extracted
(360 C), respectively. The percent of light volatiles for the
C. glomerata is 46.31 wt%. Finally, the third stage from 500 to
800 C correspond to the ulterior decomposition of the forming
bio-char. The results for proximate analysis were: 4.5 wt% moisture, 34.5 wt.% ash, 46.3 wt% volatile matter, 14.7 wt% fixed carbon,
27.8 wt% protein, 5.3 wt% lipid, and 32.4 wt% carbohydrate. Also,
the results of the BET surface area of C. glomerata was
0.15 m2 g1 and the total pore volume was 0.0028 cm3 g1.
3.1.2. Non-catalytic tests
3.1.2.1. Product distribution analysis. Pyrolysis products are gaseous, bio-oil and bio-char. The char and bio-oil were quantified
and the yields of gaseous products were calculated by difference.
Pyrolysis temperature varied from 400 to 600 C in order to determine the effect of temperature on the pyrolytic conversion of
C. glomerata into gaseous, liquid and solid products and their distributions among the total products while the reaction time and the
feed content were fixed at 30 min and 1 g, respectively. Moreover,

In order to determine the heating value of the raw C. glomerata


and its bio-char, the higher heating value (HHV) was calculated
through Eq. (5) as below (Ly et al., 2015):

HHV


Mj
f33:5  C 142:3  H  15:4  O  14:5  Ng  0:01
kg

Hydrogen selectivity was also calculated through following Eq.


(6) for performance analysis of hydrogen-rich gas production via
Pyrolysis of macroalgae.

Hydrogen Selectivity

Concentration of H2
Concentration of other gasses

Carbon liquefaction efficiency (CLE), Carbon solidification efficiency (CSE) and Carbon gasification efficiency (CGE) are mentioned below as Eqs. (7)(9) to obtain the total carbon balance
(Norouzi et al., 2016; Safari et al., 2016a)

CLE

Carbon in aqueous products


Carbon in Cladophora

CSE

Carbon in solid residues


Carbon in Cladophora

Fig. 2. Thermogravimetric and differential thermogravimetric analysis of green


macroalgae C. glomerata.

647

O. Norouzi et al. / Bioresource Technology 219 (2016) 643651

the optimum temperature at which the bio-oil portion of pyrolysis


process is maximum was determined. As shown in Table 1, by
increasing the temperature from 400 to 600 C, the gas portion
increased steadily and the char portion kept declining. The highest
gas portion obtained in this study was equaled to 28 wt% which
occurred at 600 C while The highest char yield for C. glomerata
was 44 wt% at 400 C. The highest portion of char was observed
mostly at the lower temperatures, whereas the highest yield of
gas was observed mostly at the temperatures above the 500 C.
Besides, the highest portion of bio-oil for C. glomerata was equal
to 39 wt% at the 500 C. The influence of the temperature on the
distribution of pyrolytic products from macroalgae was similar to
previous observations with lignocellulosic biomass and also other
microalgae (Li et al., 2012; Maddi et al., 2011). It was possible that
the higher ash content and lower volatile content in algal biomass
led to greater char formation but less oil production (Maddi et al.,
2011).
3.1.2.2. Gaseous product analysis. The main gaseous products produced from pyrolysis of C. glomerata were H2, CH4, CO and CO2
which are shown by their concentration in Table 1. By increasing
the temperature of pyrolysis, The H2 concentration from
C. glomerata was increased by a factor of 2.6. It is reported that
depolymerization of phenyl groups in lignin, dehydrogenation
reactions and also Protein degradation are favored at a higher temperature (Al-Rahbi et al., 2016). Hence, H2 can be produced as a
result of these reactions during pyrolysis. Moreover, it is mentioned that secondary reactions such as tar cracking are promoted
in higher temperatures which result in more H2 production. CH4
yield during pyrolysis of C. glomerata is commonly increased with
increasing temperature by the factor of 1.5. The production of
methane can be associated with the cracking of lignin and formation of methoxy groups which results in the cleaving of CAC bonds.
As seen in Table 1, the concentration of CO2 was decreased by
increasing the temperature. CO2 as the main component of the gaseous product is generally originated from carbohydrates and protein which is in agreement with the results of Yuan et al. (2015).
3.1.2.3. Liquid product analysis. The bio-oil was affected by pyrolysis temperature. The highest Bio-oil portion of 39 wt% was
obtained at the pyrolysis temperature of 500 C. Hence, this condition was selected as a representative experiment to analyze the
bio-oil composition. The bio-oil produced from non-catalytic and
catalytic pyrolysis of macroalgae at the temperature of 500 C
was collected and subjected to GCMS analysis to determine its
contained components. The identified compounds are given in
Tables 2. The nitrogenous products (29% area), were the most
prevalent in the pyrolysis of macroalgae. As expected, pyrolysis
of high-protein biomass formed a nitrogen containing liquid range
products. The acids are the second major compounds identified in
the bio-oil with the selectivity of 28.6 area% which contains
26.88 area% of acetic acid. This compound can be utilized in
microalgae cultivation with the potential of producing lipidbased biofuel (Liang et al., 2013). Phenol and its derivatives
(8.5 area%), was detected in the pyrolysis of C. glomerata. Despite

the lower amount of lignin in algae compared with agricultural


waste which tends to produce less phenol, the protein content of
algae and decomposition of oxygenate compounds results in phenols production. The phenol can be used to resin production, additives in fertilizing and pharmaceutical industries, flavoring agents
(such as glycolaldehyde) in food industries and other special chemicals (Kim, 2015).
3.1.2.4. Solid product analysis. Different analysis has been done to
determine the physical and chemical properties of bio-char for
its further application. As we know, bio-char can be used for many
applications. Porous bio-char can be utilized as a carbon activated
material for using as a catalyst support for enhancement in hydrogen production and upgrading of bio oil (Yao et al., 2016; Ren et al.,
2014). The Presence of some alkali metals in the biochar promises a
versatile material for using as a catalyst for enhancing the rate of
the reactions. According to the ICP analysis, a number of AAEMs
are as follows: 10.4 wt% Ca, 24.7% K, 7.8 wt% Mg, 14.8 wt% Na. Elemental composition of biochar was also carried out and the results
were: 45.4 wt% C, 1.1 wt% H, 3.7 wt% N, 1.7 wt% S, 8.4 wt% O and
39.7 wt% ash. The results of the BET surface area of bio-char was
42.4 m2 g1 and the total pore volume was 0.0386 cm3 g1. As
shown, the weight percentage of carbon content of C. glomerata
bio-char is increased while the hydrogen content is decreased
compared with the raw C. glomerata. This cause a higher heating
value of 17.19 MJ/kg which promises an upgraded solid fuel. On
the other hand, the BET surface area of 42.4 m2/g and the total pore
volume of 0.0386 m3/g, which are much higher than that of raw C.
glomerata, shows the high porosity of bio-char. This significant
increase in the surface area and total pore volume of feedstock
after pyrolysis are in strong agreement with previous results in
the literature (Jung et al., 2016; Hu et al., 2016). The formation of
pores in the bio-char investigated by FESEM method. According
to the FESEM analysis, numerous pores spread all over the surfaces
of the bio-char. These large pores originate from the vascular structure of the raw biomass. The thermal decomposition of volatile
matter has left the cell walls which results in the formation of
pores in the surface (Usman et al., 2015).
3.1.3. Algal bio-char as a catalyst
According to the results of non-catalytic experiments, the optimum temperature in the terms of bio-oil yield in pyrolysis of C.
glomerata was 500 C. Therefore, this pyrolysis conditions was
selected as a representative experiment to analyze product compositions and catalytic tests were performed in this condition to
observe the effect of an algal bio-char catalyst on the pyrolysis
products. Results from pyrolysis experiments are presented in
Table 3. In this study, the algal bio-char catalyst significantly influenced the product concentrations and distributions.
3.1.3.1. Effect of the algal bio-char on the product distribution. As
seen in Table 3, the addition of algal bio-char as a catalyst
improved the portion of gaseous products from 22 to 35% among
the whole product. However, the solid and liquid portion
decreased slightly to the 35% and 30%, respectively. This effect

Table 1
Effect of temperature on the product distribution and main gaseous component of the pyrolysis of C. glomerata.
Reactors temperature (C)

400
500
600

Products (wt%)

Concentration (v/v%)

Bio-oil

Bio-char

Gas

H2

CO

CO2

CH4

35
39
34

44
40
39

21
22
28

15.93
27
37.1

12.53
12.9
12.4

68.03
56.09
45.19

3.51
4.01
5.31

648

O. Norouzi et al. / Bioresource Technology 219 (2016) 643651

Table 2
Compounds identified by GCMS in C. glomerata derived bio-oil produced by catalytic and non-catalytic pyrolysis.
No.

RT (min)

Compound

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73

1.385
1.417
1.66
2.107
2.226
2.482
2.607
2.638
3.191
3.416
4.216
4.282
4.789
6.225
6.562
7.159
7.535
7.671
7.82
7.879
8.02
8.169
8.21
8.45
8.97
9.353
9.46
9.491
9.55
9.81
10.11
10.67
10.85
11.11
11.51
12.48
12.68
12.78
12.9
12.97
13.08
13.24
13.31
13.47
13.58
13.72
13.81
13.94
14.06
14.18
14.58
15.28
15.41
15.85
15.911
16.02
16.24
16.347
16.52
16.659
16.7
16.91
17.051
17.1
17.32
17.46
17.66
17.862
17.944
18.008
18.28
19.07
19.37

Trimethylamine P183
Acetone
Methylamine, N,N-dimethylAcetic acid
2-Propanone, 1-hydroxyPyrazine
2,4-Pentadienenitrile
Pyridine
Dimethylformamide P185
Phenol
Dimethylacetamide
Carbramal
Amitrol
Mepivacaine
2-Cyclopenten-1-one, 2-hydroxy-3-methylp-Cresol
Tridecane
Quinuclidine-3-ol
Perhexiline AC P859
ETHYLCYCLOPENTENOLONE
2-Pentyn-1-ol
N,N-DIMETHYLHOMOSERINE LACTONE
Pyridine
Pyrrole
Cyclopentanone
Methylpentynol P191
Pyridine, 2-methyl2,3-Anhydro-D-mannosan
Pyrazine, methyl1-Propanamine
2-Furanmethanol
1-Pentene, 2-methyl
Ethanone, 1-(2-furanyl)Pyridine, 2,4-dimethyl
2-Cyclopenten-1-one, 3-methylPhenol, 2-methylPhenol, 3-methyl2,5-Pyrrolidinedione, 1-methyl4-Hydroxy-2-hydroxyaminopyrimidine
Phenol, 2,6-dimethylMaltol
Octane, 2,7-dimethylPhenol, 2,4-dimethylPhenol, 3-ethyl1-(2-Pyrazinyl)-1-ethanol
Pyrazine, methyl-5-(1-propenyl)-, (E)Methyl Tanol
2-Propenoic acid, 2-methyl-, ethyl ester
Benzenepropanenitrile
Quinoline
5H-1-Pyrindine
1H-Indole, 5-methylNaphthalene, 2,7-dimethyl1H-Benzimidazole, 2-ethylTetradecanoic acid
Indolizine, 1,2-dimethylDecanoic acid
Pyrrolo[1,2-a]piperazine-3,6-dione
2-Formyl-1-thiocromone
Phytol
1,4-Dimethylindan-1-ol
1H-Pyrazole, 3-methyl-5-phenyl7-Thiaprotoadamantane
NEOPHYTADIENE
Pyridine, 3-(1-methyl-1H-pyrrol
2,4-Azetidinedione, 3,3-diethylTridecane, 7-methylene
Pyrrolo[1,2-a]pyrazine-1,4-dione, hexahydro-3-(2-methylpropyl)n-Hexadecanoic acid
3,9-DIAZATRICYCLO[7.3.0.0(3,7)]DODECAN-2,8-DIONE
Thiazolo [4,5-f]quinoline
2-Pentyadecanone, 6,10,14-trimethyl
9H-Fluorene, 9-methylene

Area%
Non-catalyst

Catalytic

3.08
4.4
2.15
26.88
2.09
0.67
1.6
2.1
0.68
4.23
0.59
0.94
2.56
4.66
4.52
1.68
2.77
0.5
3.29
0.1

5.72

8.35

4.325

2.25

0.76

1.01
0.44

0.91
0.64
0.5

0.96

0.51

0.67

0.53
0.37

0.56
2
0.87
4.12
1
0.12

2.88

9.91

0.14
0.31
3.01
3.778
1.209
0.791

1.99

0.85

0.88
0.67
0.38

0.36
2.532
0.33
0.97
1.35
0.28
0.75
0.38
0.83
2.2
3.79
0.84
0.34
0.69
1.11
1.11
0.97
1.99
0.88
0.61
0.59
2.23
0.78
0.4
1.95
1.96
0.53
0.62
7.72
0.37
0.51

0.35
1.41
0.49
0.41

0.36
1.08
0.52

5.68

0.97
0.35
0.3

0.37
0.714
2.241

649

O. Norouzi et al. / Bioresource Technology 219 (2016) 643651


Table 2 (continued)
No.

RT (min)

Compound

Area%

74
75
76
77
78
79
80
81
82
83
84

19.53
19.85
19.92
20.16
20.75
21.33
21.48
21.73
22.79
23.07
25.22

Tetrahydroionone
6,6-Dimethyl-cyclooct-4-enone
Hexadecanenitrile
Hexadecanoic acid, methyl ester
Z-11-Pentadecanol
2-Hydroxy-3,5,5-trimethyl-cyclohex-2-enone
9-Hydroxybutyl-3,9-diazabicyclo(3,3,1)nonane
N,N-dimethyloctanamide
2-Heptanone, O-methyloxime
Methyl 7,9-tridecadienyl ether
9-Octadeconoic acid

Non-catalyst

Catalytic

3.56
0.31
0.94
0.64
2.03
0.58
0.85
0.45
0.45
0.64
2.19

Table 3
Effect of bio-char catalyst on the product distribution and the main gaseous products of C. glomerata pyrolysis.
Tests at T = 500 C

No catalyst
Bio-char catalyst

Products (wt%)

Concentration (v/v%)

Bio-oil

Bio-char

Gas

H2

CO

CO2

CH4

39
30

40
35

22
35

27
37.08

12.90
11.8

56.09
47.09

4.01
4.03

may be associated with the more degradation of volatiles in the


presence of porous structure with the high surface area which
accelerates the reactions to form gaseous products. It is known that
the biochar has variable amounts of inorganic ash in the form of
alkali and alkaline earth metals (AAEMs) which have been suggested to catalytic activity in carbon conversion (Yildiz et al.,
2015). This catalytic effect of AAEMS was mainly because of the
increase of water molecule and also the hydrocarbon volatiles
which have been decomposed by oxygen-containing functional
groups presented in biochar. In addition to AAEMS, some transition
metals such as Iron exist inside the biomass. Despite the cracking
behavior of the Fe, the presence of it as iron oxide on the biochar
surface results in low catalytic reactivity of this metal
(Mahadevan et al., 2016).
3.1.3.2. Effect of algal bio-char on the gas composition. The effect of
the algal bio-char on the composition of pyrolysis products at the
temperature of 500 C is shown in Table 3. The addition of biochar as a catalyst increased the total gas concentration from 22%
to 37%. Noticeable differences in the concentrations of CO2 and
H2 between catalytic and non-catalytic pyrolysis were observed
but there were no significant changes in the production rate of
CO and CH4. Therefore, by adding catalyst the CO2 concentration
decreased from 56.08% to 47.09% and H2 production increased
from 27% to 37%. Enhancement of reaction activity by addition of
bio-char can be explained by results of BET and FESEM. The existence of numerous pores all over the surfaces of bio-char can provide surface reaction causing further decomposition of
complicated structure biomass (Brewer et al., 2009).

in Fig. 3, the addition of algal bio-char as a catalyst results in producing more interesting compounds. This part has been focused on
some noteworthy compounds based on the functional groups such
as acids, phenols, ketones, N-compounds, furans and alcohols, in
the terms of area%. As seen in Fig. 3, the main effect of bio-char
on the bio-oil component is the decrement in acids (28.55
23.09% for C. glomerata) which is followed by a decrease in ketones
which could be in association with the ketonization reactions, in
which acids in the pyrolysis vapor convert to ketones (Liu et al.,
2014). Moreover, the phenols were increased with the addition
of algal bio-char catalyst (8.520.76%). The furans content of the
bio-oil from C. glomerata was found to decrease from 3% to 2.1%
with the addition of bio-char. This effect can be attributed to the
presence of alkali metals (K, Na), which was reported earlier by
Mahadevan et al. (2016) this effect can be associated with decarbonylation and oligomerization reactions to produce olefins and
aromatic hydrocarbons in the presence of bio-char as a catalyst
(Mahadevan et al., 2015). These results were in agreement with
those reported in the literature (Ren et al., 2014).

3.1.3.3. Effect of the algal bio-char on the hydrogen selectivity. Hydrogen selectivity increased by the increasing of Temperature. The
highest hydrogen selectivity of 0.59 was obtained at a pyrolysis
temperature of 600 C. In addition, in the presence of algal biochar at the temperature of 500 C, H2 selectivity increased slightly
due to the catalytic effect of algal bio-char.
3.1.3.4. Effect of the algal bio-char on the bio-oil quality. The composition of bio-oil obtained from the pyrolysis of C. glomerata and its
blends with algal bio-char as a catalyst is shown in Table 2. As seen

Fig. 3. Effect of algal bio-char on the chemical composition of bio-oil.

650

O. Norouzi et al. / Bioresource Technology 219 (2016) 643651

Table 4
Carbon conversion efficiencies for catalytic and non-catalytic pyrolysis of C. glomerata.
Test#

Catalyst

Temperature

CGE

CSE

CLE

Carbon
balance

1
2
3
4

0.4 g bio-char

400
500
600
500

26.78178
26.23513
23.97984
35.38647

19.36
18.16
18.02
17.1

48.7
51
54.1
46.2

94.84177836
95.39512575
96.09983563
98.68647388

3.1.3.5. Carbon balance. Results for carbon conversion in each phase


of products has been given in Table 4. As seen, the conversion of
carbon into the liquid phase has been improved by the increase
of temperature. Also, conversion of carbon into gas phase was
enhanced by the addition of biochar due to its catalytic behavior.
Carbon balance of a total carbon conversion efficiency was closed
in 9498%.

4. Conclusion
C. glomerata as green macroalgae of Caspian Sea showed a
promising performance for production of bio-oil and hydrogen.
Non-catalytic Pyrolysis of this green macroalgae resulted in acidrich bio-oil while the addition of its bio-char as a catalyst
promoted the production of phenolics. In addition algal, bio-char
enhanced the hydrogen production and hydrogen selectivity.
FESEM, BET and ICP analysis demonstrated the porous structure
and the presence of AAEMs in the bio-char which makes it suitable
for utilization as a catalyst or catalyst support in the industry.
Acknowledgements
The authors would like to thank Dr. Ali Karimi from Research
institute of Petroleum Industry (RIPI), for his kind support for
providing facilities for characterization of materials and Sharif
University of technology, for providing the FESEM device for characterization of bio-char surface.
References
Akalin, M.K., Tekin, K., Karagoz, S., 2012. Hydrothermal liquefaction of cornelian
cherry stones for bio-oil production. Bioresour. Technol. 110, 682687. http://
dx.doi.org/10.1016/j.biortech.2012.01.136.
Ali, G., Abbas, S., Mueen Qamer, F., 2013. How effectively low carbon society
development models contribute to climate change mitigation and adaptation
action plans in Asia. Renew. Sustain. Energy Rev. 26, 632638. http://dx.doi.org/
10.1016/j.rser.2013.05.042.
Al-Rahbi, A.S., Onwudili, J.A., Williams, P.T., 2016. Thermal decomposition and
gasification of biomass pyrolysis gases using a hot bed of waste derived
pyrolysis char. Bioresour. Technol. 204, 7179. http://dx.doi.org/10.1016/j.
biortech.2015.12.016.
Brewer, C.E., Schmidt-Rohr, K., Satrio, J.A., Brown, R.C., 2009. Characterization of
biochar from fast pyrolysis and gasification systems. Environ. Prog. Sustain.
Energy 28, 386396.
Chung, I.K., Beardall, J., Mehta, S., Sahoo, D., Stojkovic, S., 2011. Using marine
macroalgae for carbon sequestration: a critical appraisal. J. Appl. Phycol. 23,
877886. http://dx.doi.org/10.1007/s10811-010-9604-9.
de la Rosa, J.M., Paneque, M., Miller, A.Z., Knicker, H., 2014. Relating physical and
chemical properties of four different biochars and their application rate to
biomass production of Lolium perenne on a Calcic Cambisol during a pot
experiment of 79 days. Sci. Total Environ. 499, 175184. http://dx.doi.org/
10.1016/j.scitotenv.2014.08.025.
Dincer, I., Rosen, M.A., 2012. Exergy: Energy, Environment and Sustainable
Development. Newnes, pp. 923 10.1007/978-3-7091-0109-4.
Guragain, Y.N., Wang, D., Vadlani, P.V., 2016. Appropriate biorefining strategies for
multiple feedstocks: critical evaluation for pretreatment methods, and
hydrolysis with high solids loading. Renew. Energy 96, 832842. http://dx.
doi.org/10.1016/j.renene.2016.04.099.
Hu, S., Jiang, L., Wang, Y., Su, S., Sun, L., Xu, B., He, L., Xiang, J., 2015. Effects of
inherent alkali and alkaline earth metallic species on biomass pyrolysis at
different temperatures. Bioresour. Technol. 192, 2330. http://dx.doi.org/
10.1016/j.biortech.2015.05.042.

Hu, Q., Yang, H., Yao, D., Zhu, D., Wang, X., Shao, J., Chen, H., 2016. The densification
of bio-char: effect of pyrolysis temperature on the qualities of pellets. Bioresour.
Technol. 200, 521527. http://dx.doi.org/10.1016/j.biortech.2015.10.077.
Jafarian, S., Tavasoli, A., Karimi, A., Norouzi, O., 2016. Steam reforming of bagasse to
hydrogen and synthesis gas using ruthenium promoted NiFe/cAl2O3 nanocatalysts. Int. J. Hydrogen Energy 18. http://dx.doi.org/10.1016/j.
ijhydene.2016.05.261.
Jones, G.A., Warner, K.J., 2016. The 21st century population-energy-climate nexus.
Energy Policy 93, 206212. http://dx.doi.org/10.1016/j.enpol.2016.02.044.
Jung, K.-W., Jeong, T.-U., Kang, H.-J., Ahn, K.-H., 2016. Characteristics of biochar
derived from marine macroalgae and fabrication of granular biochar by
entrapment in calcium-alginate beads for phosphate removal from aqueous
solution. Bioresour. Technol. 211, 108116. http://dx.doi.org/10.1016/j.
biortech.2016.03.066.
Kim, J.S., 2015. Production, separation and applications of phenolic-rich bio-oil a
review. Bioresour. Technol. 178, 9098. http://dx.doi.org/10.1016/j.
biortech.2014.08.121.
Krishnan, V., McCalley, J.D., 2016. The role of bio-renewables in national energy and
transportation systems portfolio planning for low carbon economy. Renew.
Energy 91, 207223. http://dx.doi.org/10.1016/j.renene.2016.01.052.
Li, R., Zhong, Z., Jin, B., Zheng, A., 2012. Selection of temperature for bio-oil
production from pyrolysis of algae from lake blooms. Energy Fuels 26, 2996
3002. http://dx.doi.org/10.1021/ef300180r.
Liang, Y., Zhao, X., Chi, Z., Rover, M., Johnston, P., Brown, R., Jarboe, L., Wen, Z., 2013.
Utilization of acetic acid-rich pyrolytic bio-oil by microalga Chlamydomonas
reinhardtii: reducing bio-oil toxicity and enhancing algal toxicity tolerance.
Bioresour.
Technol.
133,
500506.
http://dx.doi.org/10.1016/j.
biortech.2013.01.134.
Liu, C., Wang, H., Karim, A.M., Sun, J., Wang, Y., 2014. Catalytic fast pyrolysis of
lignocellulosic biomass. Chem. Soc. Rev. 43, 75947623. http://dx.doi.org/
10.1039/C3CS60414D.
Ly, H.V., Kim, S.S., Kim, J., Choi, J.H., Woo, H.C., 2015. Effect of acid washing on
pyrolysis of Cladophora socialis alga in microtubing reactor. Energy Convers.
Manag. 106, 260267. http://dx.doi.org/10.1016/j.enconman.2015.09.041.
Maddi, B., Viamajala, S., Varanasi, S., 2011. Comparative study of pyrolysis of algal
biomass from natural lake blooms with lignocellulosic biomass. Bioresour.
Technol. 102, 1101811026. http://dx.doi.org/10.1016/j.biortech.2011.09.055.
Mahadevan, R., Shakya, R., Neupane, S., Adhikari, S., 2015. Physical and chemical
properties and accelerated aging test of bio-oil produced from in situ catalytic
pyrolysis in a bench-scale fluidized-bed reactor. Energy Fuels 29, 841848.
http://dx.doi.org/10.1021/ef502353m.
Mahadevan, R., Adhikari, S., Shakya, R., Wang, K., Dayton, D., Lehrich, M., Taylor, S.E.,
2016. Effect of alkali and alkaline earth metals on in-situ catalytic fast pyrolysis
of lignocellulosic biomass: a microreactor study. Energy Fuels 30, 30453056.
http://dx.doi.org/10.1021/acs.energyfuels.5b02984.
Mihranyan, A., 2011. Cellulose from cladophorales green algae: from environmental
problem to high-tech composite materials. J. Appl. Polym. Sci. 119, 24492460.
http://dx.doi.org/10.1002/app.32959.
Nanda, S., Isen, J., Dalai, A.K., Kozinski, J.A., 2016. Gasification of fruit wastes and
agro-food residues in supercritical water. Energy Convers. Manag. 110, 296
306. http://dx.doi.org/10.1016/j.enconman.2015.11.060.
Norouzi, O., Safari, F., Jafarian, S., Tavasoli, A., Karimi, A., 2016. Hydrothermal
gasification performance of Enteromorpha intestinalis as an algal biomass for
hydrogen-rich gas production using Ru promoted FeNi/c-Al2O3 nanocatalysts.
Energy Convers. Manag. 210. http://dx.doi.org/10.1016/j.enconman.2016.
04.083.
Omoriyekomwan, J.E., Tahmasebi, A., Yu, J., 2016. Production of phenol-rich bio-oil
during catalytic fixed-bed and microwave pyrolysis of palm kernel shell.
Bioresour.
Technol.
207,
188196.
http://dx.doi.org/10.1016/j.
biortech.2016.02.002.
Phillips, D., Mitchell, E.J.S., Lea-Langton, A.R., Parmar, K.R., Jones, J.M., Williams, A.,
2016. The use of conservation biomass feedstocks as potential bioenergy
resources in the United Kingdom. Bioresour. Technol. 212, 271279. http://dx.
doi.org/10.1016/j.biortech.2016.04.057.
Plis, A., Lasek, J., Skawinska, A., Zuwaa, J., 2015. Thermochemical and kinetic
analysis of the pyrolysis process in Cladophora glomerata algae. J. Anal. Appl.
Pyrolysis 115, 166174. http://dx.doi.org/10.1016/j.jaap.2015.07.013.
Ren, S., Lei, H., Wang, L., Bu, Q., Chen, S., Wu, J., 2014. Hydrocarbon and hydrogenrich syngas production by biomass catalytic pyrolysis and bio-oil upgrading
over bio-char catalysts. RSC Adv. 4, 10731. http://dx.doi.org/10.1039/
c4ra00122b.
Safari, F., Salimi, M., Tavasoli, A., Ataei, A., 2016a. Non-catalytic conversion of wheat
straw, walnut shell and almond shell into hydrogen rich gas in supercritical
water media. Chin. J. Chem. Eng. http://dx.doi.org/10.1016/j.cjche.2016.03.002.
Safari, F., Tavasoli, A., Ataei, A., 2016b. Gasification of sugarcane bagasse in
supercritical water media for combined hydrogen and power production: a
novel approach. Int. J. Environ. Sci. Technol. http://dx.doi.org/10.1007/s13762016-1055-7.
Savage, P.E., 2012. Algae under pressure and in hot water. Science 338, 10391040.
http://dx.doi.org/10.1126/science.1224310.
Shoja, M., Babatabar, M.A., Tavasoli, A., Ataei, A., 2013. Production of hydrogen and
syngas via pyrolysis of bagasse in a dual bed reactor. J. Energy Chem. 22, 639
644. http://dx.doi.org/10.1016/S2095-4956(13)60084-4.
Singh, A., Nigam, P.S., Murphy, J.D., 2011. Renewable fuels from algae: an answer to
debatable land based fuels. Bioresour. Technol. 102, 1016. http://dx.doi.org/
10.1016/j.biortech.2010.06.032.

O. Norouzi et al. / Bioresource Technology 219 (2016) 643651


Sluiter, B.H.A., Ruiz, R., Scarlata, C., Sluiter, J., Templeton, D., 2008. Determination of
Ashin Biomass, Technical Report NREL/TP-510-42622, National Renewable
Energy Laboratory.
Soltani, S., Ebrahimzadeh, M.A., Khoshrooei, R., Rahmani, Z., 2014. Evaluation of
antibacterial activities in Cladophora glomerata and Enteromorpha intestinalis.
Int. J. Mol. Clin. Microbiol. 1, 371376.
Usman, A.R.A., Abduljabbar, A., Vithanage, M., Ok, Y.S., Ahmad, M., Ahmad, M.,
Elfaki, J., Abdulazeem, S.S., Al-Wabel, M.I., 2015. Biochar production from date
palm waste: charring temperature induced changes in composition and surface
chemistry. J. Anal. Appl. Pyrolysis 115, 392400. http://dx.doi.org/10.1016/
j.jaap.2015.08.016.
Vassilev, S.V., Vassileva, C.G., 2016. Composition, properties and challenges of algae
biomass for biofuel application: an overview. Fuel 181, 133. http://dx.doi.org/
10.1016/j.fuel.2016.04.106.

651

Yao, D., Hu, Q., Wang, D., Yang, H., Wu, C., Wang, X., Chen, H., 2016. Hydrogen
production from biomass gasification using biochar as a catalyst/support.
Bioresour. Technol. http://dx.doi.org/10.1016/j.biortech.2016.05.011.
Yildiz, G., Ronsse, F., Venderbosch, R., van Duren, R., Kersten, S.R.A., Prins, W., 2015.
Effect of biomass ash in catalytic fast pyrolysis of pine wood. Appl. Catal., B:
Environ. 168169, 203211. http://dx.doi.org/10.1016/j.apcatb.2014.12.044.
Yuan, T., Tahmasebi, A., Yu, J., 2015. Comparative study on pyrolysis of
lignocellulosic and algal biomass using a thermogravimetric and a fixed-bed
reactor. Bioresour. Technol. 175, 333341. http://dx.doi.org/10.1016/j.
biortech.2014.10.108.
Zhao, B., Wang, X., Yang, X., 2015. Co-pyrolysis characteristics of microalgae
Isochrysis and Chlorella: kinetics, biocrude yield and interaction. Bioresour.
Technol. 198, 332339. http://dx.doi.org/10.1016/j.biortech.2015.09.021.

You might also like