Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Energy Conversion and Management 95 (2015) 1019

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

Optimization of renewable levulinic acid production from glucose


conversion catalyzed by Fe/HY zeolite catalyst in aqueous medium
Nur Aainaa Syahirah Ramli, Nor Aishah Saidina Amin
Chemical Reaction Engineering Group (CREG), Energy Research Alliance, Faculty of Chemical Engineering, Universiti Teknologi Malaysia, 81310 UTM Johor, Johor Bahru, Malaysia

a r t i c l e

i n f o

Article history:
Received 10 December 2014
Accepted 5 February 2015
Available online 21 February 2015
Keywords:
Levulinic acid
Optimization
Glucose
Oil palm fronds
Fe/HY zeolite
Response surface methodology

a b s t r a c t
Levulinic acid (LA) is a versatile chemical with numerous applications. In this study, the conversions of
glucose and oil palm fronds (OPF) to LA have been conducted over 10% Fe/HY zeolite catalyst. The optimization of LA yield from glucose conversion using BoxBehnken design and response surface methodology reported 61.8% yield, which can be achieved at temperature 173.4 C, reaction time 3.3 h, 0.93 g of
glucose and 0.89 g 10% Fe/HY zeolite. The LA yield from OPF conversion conducted at the optimum conditions was 17.6% with 54.8% process efciency. It was also observed that Fe leaching from 10% Fe/HY
zeolite was insignicant and recycled 10% Fe/HY zeolite gave sufcient performance for ve successive
cycles. This study emphasizes the potential of Fe/HY zeolite catalyst for catalytic conversion of lignocellulosic biomass to LA.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
Primary energy sources such as petroleum, natural gas and coal
are all non-renewable, and depending on these sources for energy
generation is not sustainable. The depleting primary energy
sources and increasing greenhouse gas emissions have driven
worldwide initiatives to develop clean technologies for chemical
transformations from renewable and sustainable feedstocks. Biomass has received signicant attention for its ability to produce
fuels and chemicals. In Malaysia, the biomass residues culminate
mostly from the oil palm industry. Oil palm fronds (OPF) are the
most abundant oil palm wastes generated by the mills. OPF contains large amount of cellulose, the polymer of glucose, which
can be further converted downstream to other value added products [1]. This study presents the report of OPF conversion to
levulinic acid (LA), which is one of the top twelve building blocks
derived from biomass feedstock [2]. The production of value added
products from OPF in a biorenery may provide a better method to
manage oil palm wastes for saving the environment and improving
the sustainable development of the oil palm industry.
LA is a versatile building block for the production of various platform chemicals. Besides fuel additives, LA can also be used in
manufacturing biodegradable herbicides, solvents, food avoring
agents, fragrance, pharmaceutical compounds, and resins [3]. LA
can be synthesized from various feedstocks including glucose,
Corresponding author. Tel.: +60 75535579; fax: +60 75588166.
E-mail address: noraishah@cheme.utm.my (N.A.S. Amin).
http://dx.doi.org/10.1016/j.enconman.2015.02.013
0196-8904/ 2015 Elsevier Ltd. All rights reserved.

starch and lignocellulosic materials [49]. Generally, in the conversion of glucose to LA, glucose isomerizes to fructose and dehydrates
to form the intermediate product, 5-hydroxymethylfurfural (5HMF) [1013], which will then catalytically rehydrate to form LA
and formic acid. Due to the availability of cheap biomass, the direct
transformation of biomass to LA has gained extensive attention
from many researchers worldwide [58,1416].
Previously, numerous studies have reported the production of
LA using mineral acids such as HCl, H2SO4, and formic acid as catalysts [5,17,18]. A homogeneous catalytic system has been applied
in the rst biorenery in Caserta, Italy for the synthesis of LA from
biomass through a process developed by Bione Renewables
[19,20]. Although homogeneous catalytic reactions were effective
for LA production, the use of mineral acids has caused serious environmental pollution [4].
From an economical and environmental point of view, heterogeneous solid acid catalysts have attracted substantial interests
since these solid catalysts can overcome the problems associated
with homogeneous catalytic system [21]. Many efforts have been
attempted to employ recyclable and reusable solid acid catalysts
for the production of LA. Some of the catalysts were zeolites, metal
salts, ion-exchange resins, niobium phosphate, and solid super acid
[4,79,22,23]. It was demonstrated that zeolite has the potential to
be used in glucose dehydration reaction since the pore structure
could exert signicant inuence on the reaction [9,24,25]. Metal
halides have also been used for LA production from cellulose
[22]. Metal halides are expected to demonstrate superior catalytic
activity for glucose dehydration due to its acidity. It also has the

N.A.S. Ramli, N.A.S. Amin / Energy Conversion and Management 95 (2015) 1019

additional benet of being easily separated from reaction products


by supporting the metal halides on a carrier such as zeolite [26]. In
earlier research, CrCl2 or CrCl3 were mostly involved as catalyst for
catalytic conversion of carbohydrates [22,2731]. However, high
price, toxicity and environmental pollution caused by CrCl2 and
CrCl3 have necessitated the search for a non toxic and low cost catalyst. Instead of using Cr, other metal chlorides can be considered
Due to the low cost and non toxic properties of FeCl3 [32], it is
envisaged that this extensively available compound is able to perform as an eco-friendly catalyst for conversion of carbohydrates to
LA.
Optimization process is a crucial step in determining suitable
and economical conditions to enhance process efciency. Response
surface methodology (RSM) is managed by certain rules where the
relationship between the factors can be used to optimize a desired
output [33]. The application of RSM can decrease the number of
experimental runs needed and reduce the cost and time consumed
accordingly. In previous studies, RSM have been employed to optimize LA yield from glucose and biomass conversions [5,15,34,35].
The important variables which possibly affect the LA yield include
reaction temperature, reaction time, feedstock loading, and catalyst loading [8,15,23,34,36].
Recently, CrCl3 and HY zeolite were combined for the production of LA from glucose conversion [34,37]. Herein, Fe/HY zeolite,
prepared by combination of FeCl3 and HY zeolite through wet
impregnation method, was explored for the optimization of LA
yield from catalytic conversion of glucose. The activity and characterization of different FeCl3 loading on HY zeolite have been
evaluated for glucose dehydration reaction [38]. The Fe/HY zeolite
catalyst has previously been tested for reducing sugar production
from the hydrolysis of cellulose and oil palm biomass in ionic liquid for further production of LA [39]. The combination of Fe/HY
zeolite catalyst and ionic liquid has also been tested for direct conversion of OPF to LA [40]. The Fe/HY zeolite-ionic liquid catalytic
system was effective for LA production. However, the exorbitant
cost of ionic liquid restricted its extensive application. This poses
the challenge to investigate the catalytic activity of Fe/HY zeolite
for the conversion of biomass in aqueous medium.
Thus, the present work demonstrates the optimization of LA
yield from glucose conversion catalyzed by Fe/HY zeolite catalyst
in aqueous medium. Initially, Fe/HY zeolite catalysts with different
FeCl3 loadings (5, 10 and 15 wt%) were screened. The effect of
Fe/HY zeolite properties on the production of LA from glucose
was examined and briey discussed. Next, the optimization of LA
yield from glucose conversion over the best Fe/HY zeolite catalyst
using Box Behnken design under RSM was performed. The study is
essential for determining the optimum conditions for better utilization of Fe/HY zeolite catalyst in an industrial scale process.
The effect of various process conditions; reaction temperature,
reaction time, glucose loading, and Fe/HY zeolite loading on LA
yield as the response was evaluated. The signicance of process
variables and the interaction between variables on the process
were also examined using RSM. Subsequently, the potential of
Fe/HY zeolite as catalyst was evaluated on OPF conversion to LA
at the optimum condition obtained from the glucose conversion
reaction. In addition, leaching of Fe ions from Fe/HY zeolite catalyst
and the reusability of Fe/HY zeolite catalyst for LA production from
glucose and OPF conversions were also examined.

zeolite (SiO2/Al2O3 = 5) was obtained from Zeolyst International,


U.S.A. Ammonium chloride (NH4Cl) and sulfuric acid (H2SO4) were
supplied by Qrec, New Zealand. OPF was supplied by Malaysia
Palm Oil Board (MPOB), Kuala Lumpur, Malaysia. The whole OPF,
consisted of leaets and petiole, were dried and grinded to small
size particles (less than 5 mm).
2.2. Catalyst preparation and characterization
HY zeolite was prepared by exchanging the ion from NaY zeolite
with NH4Cl. NaY zeolite was contacted with 2 M NH4Cl with stirring (250 rpm) at room temperature for 2 h. The precipitate was
then washed with distilled water, followed by drying overnight
at 120 C. The material was calcined at 400 C for 5 h resulting in
the HY zeolite. The Fe/HY zeolite was prepared using wetness
impregnation method [38]. FeCl3 solution and HY zeolite with certain weight ratio of FeCl3 to HY zeolite (5%, 10%, 15%) were mixed
and stirred (250 rpm) at room temperature for 2 h. Then, the mixture was dried in the oven overnight at 120 C. Finally, the Fe/HY
zeolite catalyst was calcined at 400 C for 5 h.
The properties of Fe/HY zeolite and HY zeolite catalysts were
examined using several characterization methods. EDX analysis
was conducted to conrm the weight percentage of Fe on Fe/HY
zeolite samples. N2 physisorption method was applied to determine the surface area of the catalysts using a Micromeritics
ASAP2020 analyzer. The acidity of the catalysts was evaluated
using temperature programmed desorption of ammonia
(NH3-TPD, TPDRO 1100 series, Thermo Finnigan). The Brnsted to
Lewis acid sites ratios were determined by Perkin Elmer FTIR spectroscopy with pyridine as a probe molecule. The pyridine adsorption was performed at 250 C for 1 h. Besides, FTIR, XRD, FESEM,
N2 physisorption, and NH3-TPD analyses were conducted to examine the regenerated catalyst. The FTIR spectra were recorded using
Perkin Elmer at 4004000 cm1 and the FESEM images were
obtained using FESEM, Hitachi SU8020. Meanwhile, the XRD was
employed using Bruker D8 Advance diffractometer system (Cu
Ka radiation).
2.3. Catalytic tests
All experiments were carried out in a 100 ml high pressure
reactor. The reactor was loaded with predetermined amount of
glucose and Fe/HY zeolite catalyst in distilled water (50 ml) and
heated at the specied temperature and constant stirring speed
(200 rpm). After the reaction was completed, the product was let
cooled to room temperature and ltered prior to product analysis.
The concentration of LA in the liquid product was determined by
HPLC (Perkin Elmer Series 200) using column, Hi Plex H; ow rate,
0.6 ml/min; mobile phase, 5 mM H2SO4; detector, UV210 nm;
retention time, 60 min; column temperature, 60 C. The same steps
were applied to OPF hydrolysis. The LA yield from glucose and OPF,
theoretical LA yield from OPF, and process efciency of OPF hydrolysis are calculated according to Eqs. (1)(3). Other water soluble
products from glucose conversion reaction (5-HMF, formic acid,
furfural) were detected using HPLC (Perkin Elmer Series 200) under
the same conditions as LA determination. The products yield is calculated according to Eq. (4):

Final LA amount
 100%
Initial feedstock amount
Cellulose amount  0:71
Theoretical LA yield %
 100%
Oil palm frond amount

LA yield %
2. Materials and methods
2.1. Materials
Iron (III) chloride (FeCl3), glucose, LA, 5-HMF, formic acid, and
furfural were purchased from Merck, Germany. NaY faujasite type

11

1
2

where the value of 0.71 is equal to the molecular weight of LA divided by the molecular weight of cellulose (MW of LA, C5H8O3 = 116,
MW of cellulose, C6H10O5 = 162).

N.A.S. Ramli, N.A.S. Amin / Energy Conversion and Management 95 (2015) 1019

Efficiency %

LA yield
 100%
Theoretical LA yield

where efciency refers to the efciency of OPF conversion to LA


based on the cellulose content.

Product yield %

Final product amount


 100%
Initial glucose amount

2.4. Experimental design


In this study, the BoxBehnken design (BBD) was used to design
the experiments with four variables. The variables; reaction temperature (x1), reaction time (x2), glucose loading (x3), and 10%
Fe/HY zeolite loading (x4), at three different levels; low, medium,
high, and coded as 1, 0, +1, respectively (Table 1). The statistical
analysis of the response was carried out using Statsoft Statistica
software version 8.0. The mathematical model for LA yield was
tted to second order polynomial model as in Eq. (5):

Y i bo b1 x1 b2 x2 b3 x3 b4 x4 b11 x21 b22 x22 b33 x23


b44 x24 b12 x1 x2 b13 x1 x3 b14 x1 x4 b23 x2 x3 b24 x2 x4
b34 x3 x4

where Yi is dependent variable (LA yield); x1, x2, x3, and x4 are the
independent variables; bo is the regression coefcient; b1, b2, b3,
and b4 are the linear coefcients; b11, b22, b33, and b44 are the
quadratic coefcients; b12, b13, b23, b34, b24, and b14 are the second
order interaction coefcients.
2.5. OPF characterization
The chemical compositions of OPF were determined by thermal
gravimetric analysis (TGA) using a NETZSCH STA 449F3 instrument. The OPF sample was heated with heating rate of 10 C/min
under N2 ow, from 30 to 700 C. TGA method has been proven
as a method to determine biomass compositions [20,41]. The characterization of sugars and ash content in the OPF were in accordance with the standard laboratory analytical procedures
provided by the National Renewable Energy Laboratory (NREL)
[42].
3. Results and discussions
3.1. Preliminary testing and Fe/HY zeolite characterization
The rst step in this study was screening among the three Fe
modied HY zeolite based catalysts with different FeCl3 loading;
5%, 10% and 15%, through glucose conversion to LA. These catalysts
are referred as 5% Fe/HY zeolite, 10% Fe/HY zeolite, and 15% Fe/HY
zeolite. The operation conditions were 180 C, 3 h, 1 g of glucose,
1 g of catalyst, and 50 ml of distilled water. As mentioned earlier,
this paper deals with the optimization of LA from glucose conversion using RSM and the application of Fe/HY zeolite for OPF conversion to LA while our previous paper [38] is more focused and
detailed on the catalysts characterization and performance. As it

Table 1
Experimental range and levels for the independent variables.
Factors

Reaction temperature (C)


Reaction time (h)
Glucose loading (g)
10% Fe/HY zeolite loading (g)

Symbol

was discussed in previous study, the amount and type of acid sites
together with surface area and porosity inuenced the glucose
conversion to LA [38]. The Fe/HY zeolite catalyst with a large surface area, high concentration of active acid sites, and appropriate
ratio of Brnsted to Lewis acids seemed suitable for LA production.
As comparison, glucose conversion to LA was also tested using the
parent catalysts; FeCl3 and HY zeolite. The results of glucose conversion to LA for each catalyst are depicted in Fig. 1. Among the
catalysts tested, the highest LA yield was observed over 10%
Fe/HY zeolite catalyst. Thus, 10% Fe/HY zeolite catalyst was used
to determine the optimum conditions to give optimum LA yield
from glucose, and subsequent application of the catalyst on LA production from OPF.
The EDX results showed the appearance of Fe, Si, Al, and O for
all Fe/HY zeolites, and Si, Al, and O for HY zeolite (Table S1). The
results have conrmed the weight percentage of Fe in the Fe/HY
zeolite samples. Catalyst characterizations using N2 physisorption,
NH3-TPD, and pyridine-FTIR were performed to summarize why
10% Fe/HY zeolite was chosen for the optimization study and OPF
conversion. The N2 physisorption analysis revealed that the specic surface area of HY zeolite was reduced after the impregnation
process (Table S2). The NH3-TPD proles of HY zeolite and Fe/HY
zeolites demonstrate several desorption peaks existed in all catalysts (Fig. S1). The total acidity of the zeolite catalyst increased
after being impregnated with FeCl3 (Table S2), where 5% Fe/HY
zeolite has the highest total acidity amongst all the catalysts tested. The total acidity of 10% Fe/HY zeolite and 15% Fe/HY zeolite
were lower than 5% Fe/HY catalyst although the FeCl3 loading
was higher; owing to the presence of Fe2O3 covering the acid sites
on the HY zeolite surface [38]. However, the number of acid
sites (Table S2) are in accordance to the following order: 10%
Fe/HY > 5% Fe/HY > 15% Fe/HY > HY. This suggests that 10% Fe/HY
zeolite contained more reactive active sites compared to the other
catalysts. The number of acid sites were calculated by dividing the
total acidity with specic surface area.
Meanwhile, the pyridine-FTIR was used to distinguish between
Brnsted and Lewis acid sites of the catalysts. From the spectra,
Fe/HY zeolite catalysts exhibited higher Lewis acid sites compared
to Brnsted acid sites (Fig. S2). The decrease of Brnsted to Lewis
acid sites ratio of Fe/HY zeolites indicate that the impregnation
of FeCl3 on HY zeolite increased the Lewis acid sites on the catalyst
surface (Table S2) [38]. In glucose conversion to LA, Lewis acid sites
has been reported to favor the isomerization of glucose into fructose, whereas the combination of Brnsted and Lewis acidity favor
the dehydration/rehydration reactions [23,4345]. However, the
increase in Lewis acid sites increased the formation rate of
unwanted by-product, humins and consequently decrease the LA
formation [43]. Hence, an appropriate Brnsted to Lewis acid sites
ratio are essential for the production of LA.

60

LA yield (%)

12

40

20

Range and level

Uncoded

Coded

1

+1

X1
X2
X3
X4

x1
x2
x3
x4

140
2
0.5
0.5

170
3
1
1

200
4
1.5
1.5

0
5% Fe/HY
zeolite

10% Fe/HY
zeolite

15% Fe/HY
zeolite

FeCl3

HY zeolite

Catalyst
Fig. 1. LA obtained from glucose conversion using different catalysts.

13

N.A.S. Ramli, N.A.S. Amin / Energy Conversion and Management 95 (2015) 1019

The trend in LA prole implies that number of acid sites together with Brnsted and Lewis acid sites present in the catalyst surface affect the LA production. From the preliminary testing, the
LA yield over 15% Fe/HY zeolite was lower compared to 5% Fe/HY
and 10% Fe/HY zeolite catalysts. As aforementioned, this might
be due to the highest Lewis acid sites that promote the formation
of humins. Among all the tested catalysts, HY zeolite exhibited the
highest Brnsted acid sites, and yet the LA yield was lower compared to the LA yields produced over Fe/HY zeolite catalysts. The
trend is attributed to the smallest number of acid sites displayed
by the HY zeolite. Thus, it is surmised that glucose conversion to
LA over Fe/HY zeolite is predominantly inuenced by the catalyst
acidity and active catalytic sites derived from HY zeolite.
The effects of reaction time and catalyst loading have been
briey inspected on glucose conversion over 10% Fe/HY zeolite
[38]. In order to produce LA, water is need in the reaction system,
otherwise 5-HMF may be converted to humins [18]. The effect of
water to glucose by weight ratio on LA yield was examined.
According to Fig. S3, no signicant effect of water to glucose ratio
has been detected on LA yield. Thus, constant amount of water
(50 ml) was used through out the reactions.
3.2. RSM study for LA production from glucose
3.2.1. Model analysis
According to the preliminary testing the best catalyst was 10%
Fe/HY zeolite. Thus, 10% Fe/HY zeolite was used in the optimization
of LA yield from glucose conversion. The process was investigated
according to BBD (27 batch experiments) and the results are shown
in Table 3. The second order polynomial model for LA yield from
glucose conversion using 10% Fe/HY zeolite catalyst is as in Eq. (6):

Y i 268:933 2:555x1 41:133x2 30:533x3 63:1x4


 0:007x21  5x22  15:95x23  9:65x24  0:005x1 x2
0:057x1 x3  0:183x1 x4  3:55x2 x3  4:55x2 x4 x3 x4

The model indicated the sufciency between the observed and


predicted response, where the coefcient of determination (R2)
was close to 1. The values of R2 for LA model was 0.94 as shown
in Fig. 2a. The results indicated that 94% of the variability in the
response can be explained by the model. The high R2 value indicates the model is competent to give a good estimate of the
response within the process condition range [34]. Fig. 2a compares
the predicted and observed LA yield. The results specify that the
observed LA yield was scattered relatively near to the straight line,
and the sufcient correlation between these values was observed.
The analysis of variance (ANOVA) as tabulated in Table 3 was
used to check the calculated F value and comparing it with the
tabulated F-value. The tabulated F-value was used at a high condence level; 95%, in order to obtain a good prediction model. Generally, the calculated F-value should be greater than the tabulated
F-value to reject the null hypothesis. The calculated F-values for LA
yield model is 13.34 which is higher than the tabulated F-value by
rejecting the null hypothesis at the 0.05 level of signicance.
The signicance of each coefcient and the pattern of interactions between the factors, were determined by t-value and p-value
as shown in the Pareto chart (Fig. 2b). The greater the t-value and
the smaller the p-value, the more signicant the corresponding
coefcient is [5,46]. As illustrated in Fig. 2b, with smaller p-value
and larger t-value, the quadratic term of all variables; reaction
temperature (x21), reaction time (x22), glucose loading (x23), and 10%
Fe/HY zeolite loading (x24) gave a remarkable effect on the LA yield
compared to the linear terms. This suggests that increasing each
individual parameter under the experimental conditions will
increase the LA yield until it reached its optimum value. Besides,
there are also linear term interactions between reaction temperature and 10% Fe/HY zeolite loading (x1x4), reaction time and
10% Fe/HY zeolite loading (x2x4), and reaction time and glucose
loading (x2x3). Most of the signicant coefcients are convincing
at 99% signicant level by rejecting the null hypothesis at 1% level,
while some of the coefcients are signicant at 95% signicant level. With p-value higher than 0.05, it is conrmed that there are less

Table 2
Experimental data set for LA yield from glucose.
Run

Variables

Response

x1

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

x2

x3

x4

Y1

Reaction temperature (C)

Level

Reaction time (h)

Level

Glucose loading (g)

Level

10% Fe/HY zeolite loading (g)

Level

LA yield (%)

140.0
140.0
200.0
200.0
170.0
170.0
170.0
170.0
140.0
140.0
200.0
200.0
170.0
170.0
170.0
170.0
140.0
140.0
200.0
200.0
170.0
170.0
170.0
170.0
170.0
170.0
170.0

1
1
+1
+1
0
0
0
0
1
1
+1
+1
0
0
0
0
1
1
+1
+1
0
0
0
0
0
0
0

2.0
4.0
2.0
4.0
3.0
3.0
3.0
3.0
3.0
3.0
3.0
3.0
2.0
2.0
4.0
4.0
3.0
3.0
3.0
3.0
2.0
2.0
4.0
4.0
3.0
3.0
3.0

1
+1
1
+1
0
0
0
0
0
0
0
0
1
1
+1
+1
0
0
0
0
1
1
+1
+1
0
0
0

1.0
1.0
1.0
1.0
0.5
0.5
1.5
1.5
1.0
1.0
1.0
1.0
0.5
1.5
0.5
1.5
0.5
1.5
0.5
1.5
1.0
1.0
1.0
1.0
1.0
1.0
1.0

0
0
0
0
1
1
+1
+1
0
0
0
0
1
+1
1
+1
1
+1
1
+1
0
0
0
0
0
0
0

1.0
1.0
1.0
1.0
0.5
1.5
0.5
1.5
0.5
1.5
0.5
1.5
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
0.5
1.5
0.5
1.5
1.0
1.0
1.0

0
0
0
0
1
+1
1
+1
1
+1
1
+1
0
0
0
0
0
0
0
0
1
+1
1
+1
0
0
0

46.8
51.4
48.2
52.2
55.8
54.9
53.7
53.8
50.2
54.6
58.7
52.1
50.9
53.2
57.6
52.8
53.2
50.2
53.4
53.8
49.7
55.7
59.9
56.8
62.1
62.3
61.9

14

N.A.S. Ramli, N.A.S. Amin / Energy Conversion and Management 95 (2015) 1019

Predicted levulinic acid yield (%)

64
R 2 =0.94

60

56

52

(a)

48
44

48

52

56

60

64

Observed levulinic acid yield (%)


p -value
x12

0.000001

x22

0.000005

x32
x2

0.000048

x4

9.756
7.743
6.175

0.000274

5.072

0.002843

3.736

x1x 4

0.003103

3.688

x2x 4

0.010065

x2x 3

0.034732

x1

0.038565

x3

0.134107

x1x 3

0.276550

x3x 4

0.743203

x1x 2

0.843937

x4

0.984874

3.051
2.380
2.323
1.607
1.140
0.335

(b)

0.201
0.019

p-value = 0.05

Standardized Effect Estimate (Absolute Value)


Fig. 2. (a) The coefcient of determination and predicted versus observed LA yield.
(b) Pareto chart of LA yield model.

interactions between reaction temperature and glucose loading


(x1x3), glucose loading and 10% Fe/HY zeolite loading (x3x4), and
reaction temperature and reaction time (x1x2) on the LA yield from
glucose conversion.
3.2.2. Variables effect on the response
The three-dimensional (3D) response surface plots and contour
plots which interpreted the interaction effects between two variables and the LA yield are presented in Fig. 3. These plots were
drawn by varying two variables while the other variable was maintained at zero level. The interaction effects were considered within
the range of variables. The smallest ellipse in the contour plot indicates the optimum values of variables at the maximum predicted
response [5].
From Fig. 3ac, the interactions of reaction temperature with
reaction time, glucose loading, and 10% Fe/HY zeolite loading,
respectively, were observed attesting quadratic effect of reaction
temperature on LA yield. Fig. 3a revealed that LA yield linearly
increased with increasing reaction temperature and time up to
the optimum conditions. The circular nature of contour plot in
Fig. 3a shows less interaction between these two variables. As
mentioned before, it is known that elevated temperature can

Table 3
Analysis of variance (ANOVA) for quadratic model of LA yield.
Sources

Sums of
square

Degree of
freedom

Mean
square

Fvalue

F0.05,14,12

Regression
Residual
Total

415.29
26.69
441.98

14
12
26

29.66
2.24

13.34

2.64

contribute to the acceleration of reaction rate and conversion efciency. At higher temperature, atoms give up or receive electrons
more easily, thus increasing the chemical reaction rate [5,14].
The interaction between reaction temperature and glucose loading
on LA yield at 3 h and 1 g 10% Fe/HY zeolite is illustrated in Fig. 3b.
The contour plot signies small interaction between reaction temperature and glucose loading. Besides, the plot implies that reaction temperature seemed to give more effect on the reaction
compared to glucose loading.
Fig. 3c exhibits that the interaction between reaction temperature and 10% Fe/HY zeolite loading is relatively signicant as
indicated by the surface conned in the smallest ellipse in the contour diagram as conrmed by the p-value = 0.003103. With
increasing 10% Fe/HY zeolite loading, the LA yield also increased
due to the higher number of active sites available in the reaction
system. After the 10% Fe/HY zeolite loading met the optimum
requirement, higher reaction temperature resulted in lower LA
yield. In this process, conversion of glucose produced LA as the
main soluble product. In addition, relatively low product concentrations of 5-HMF, formic acid, and furfural were also detected.
The carbonaceous residue was formed at higher temperature, prolonged reaction time, higher glucose loading, and higher 10% Fe/HY
zeolite loading, inferring unwanted side reactions also increased at
the same time. The same results have also been observed previously [22,34]. From the 3D response surface plots, it is conrmed that
reaction temperature exceeding the optimal is unfavorable for
increasing LA yield.
Fig. 3d and e demonstrate the elliptical nature of the contour
plots for reaction time with glucose and 10% Fe/HY zeolite loading. The contour plots indicated the signicant interactions of
reaction time with glucose and 10% Fe/HY zeolite loading on LA
production. As shown in Fig. 3d and e, LA yield increased with
longer reaction time at lower glucose and 10% Fe/HY zeolite loading. In the limited amount of acid catalyst, prolonged reaction
time was needed for complete conversion of glucose to LA. The
same trend has also been reported previously for homogeneous
acid catalyst [15,47]. With the enrichment of glucose loading,
LA yield decreased as a result of limited active sites of the catalyst
[22]. On the other hand, lower LA yield at higher 10% Fe/HY
zeolite loading and prolonged reaction time was probably due
to the excess active sites of the catalyst that boost up the hydrolysis process, and at the same time promote undesired side reactions. Yet, Fang and Hanna have suggested that increasing the
acid concentration and reaction time may improved the conversion of feedstock even at higher loading [14]. These results suggested that if the reaction is continued for too long, the LA
yield would decline as occurrence of undesired side reactions
and accumulation of undesired byproducts.
At 170 C and 3 h reaction conditions, the circular nature of the
contour plot (Fig. 3f) can be observed. The contour plot indicated
less interactions between glucose and 10% Fe/HY zeolite loading
on the LA yield. The LA yield increased to the maximum and then
decreased after overloading glucose and 10% Fe/HY zeolite in the
reaction system [22]. Higher glucose loading would increase the
probability of the reactive compounds such as glucose, fructose,
and 5-HMF collided with each other and caused cross polymerization to produce undesired byproducts, such as humins. Similarly,
higher 10% Fe/HY zeolite loading decreased the LA yield, possibly
as a consequence of the unwanted side reactions. Besides, higher
glucose and catalyst loading might resulted in the reduction of
mixing effect thus limiting more glucose to join in reaction at the
same time. Thus, no more catalyst should be loaded once the
amount of catalyst meets its necessity. This statement has been
supported before [34,48]. Hence, the optimal 10% Fe/HY zeolite
and glucose loading for maximum LA yield have to be scrutinized
based on economic prospect of the process.

15

N.A.S. Ramli, N.A.S. Amin / Energy Conversion and Management 95 (2015) 1019

(b)
Levulinic acid yield (%)

Levulinic acid yield (%)

(a)
65
60
55
50
45
40

4.0

56
52
48
44
40

210

3.6

ac

2.8

2.4
2.0

(h

130

tio

ac

Re

era

p
tem

150

(d)

64

64

Levulinic acid yield (%)

(c)
60
56
52
48
44
40

1.4

Ca

ta

lys

1.0

t lo

ing

150

0.6

(g

130

1.4

lys

3.4

1.0

t lo

4.2

3.8

1.2
3.0

0.8

ad

ing

2.6

0.6

(g

2.2
1.8

h)

e(

im
nt

tio

ac

Re

3.0

0.8

din

2.6

0.6

g(

Levulinic acid yield (%)

44

ta

3.4

1.0

oa

64

Ca

3.8

os

el

64

48

ac

Re

1.2

uc

(f)

52

t
ion

1.4

Gl

(e)

56

pe

em

130

44

g)

60

)
(o C

48

Re

150

re
atu

52

em

nt

ctio

170

56

)
190
(o C
e
r
tu
era

170

0.8

ad

190

60

210
1.2

210

uc 1.2
os
1.0
el
oa 0.8
din
0.6
g(
g)

tur

170

tim

1.4

Gl

o C)
e(

190

3.2

tio

Levulinic acid yield (%)

60

35

Re

Levulinic acid yield (%)

64

2.2

h)

e(

im
nt

tio

ac

Re

1.8

4.2

60
56
52
48
1.4

Ca

1.4

1.2

ta

lys

1.2

1.0

t lo

1.0

0.8

ad

ing

0.8

0.6

(g

0.4

0.6

se

ing

1.6

(g)

d
loa

co

Glu

Fig. 3. Response tted surface area plots of LA yield versus different variables.

3.2.3. Optimization study


Based on the model analysis, the LA yield is forecasted at optimum conditions in order to achieve high LA yield from glucose
conversion. The response surface analysis indicates that the

predicted optimum LA yield from glucose is 62.6% which can be


achieved from reaction conducted at 173.4 C, 3.3 h reaction time,
using 0.93 g glucose, and 0.89 g 10% Fe/HY zeolite. Further test was
conducted at the optimum conditions for the conrmation of the

16

N.A.S. Ramli, N.A.S. Amin / Energy Conversion and Management 95 (2015) 1019

O
O
Si
Fe
O
O
OH

OH
O
HO
OH

+ Fe/HY

OH

OH

OH
O

OH
OH

OH

OH

OH

+ Fe/HY

OH

OH

HO

OH

- Fe/HY
HO

OH

HO

OH

HO

- H2O

tautomerization
OH

OH

OH

OH

OH

OH

OH

- 2H2O
HO

O
O

+ 2H2O

OH

- HCHO

OH

O
O

10
Fig. 4. Proposed reaction mechanism of LA production from glucose over Fe/HY zeolite.

3.3. Proposed mechanism of LA production from glucose using Fe/HY


zeolite
A possible mechanism that might take place in LA production
from glucose over Fe/HY zeolite catalyst is illustrated in Fig. 4.
The reaction was initiated by isomerization of glucose to fructose.
The active sites for the isomerization of glucose using Fe/HY zeolite
are Fe atoms incorporated in the zeolite framework. Glucose (1)
was converted into open chain form (2) through complex formation with Fe/HY zeolite catalyst. Next, the open chain glucose
was combined with Fe/HY to form an enolate intermediate (3).
The enolate was converted into fructose (4) by the release of
Fe/HY zeolite. Similarly, few studies have reported that fructose
was produced from glucose through an enediol intermediate
accelerated by the catalyst used [50,51]. The isomerization reaction of glucose into fructose might also take place by hydrogen
transfer through a hydride shift between carbonyl containing C1

and hydroxyl bearing C2 [52]. The hydride shift is due to the presence of Fe metal in the zeolite framework.
The enolate formation would allow the conversion of glucose
into fructose, followed by dehydration reaction to produce 5HMF. A detailed mechanism of fructose dehydration to 5-HMF
has been explained by Caratzoulas and Vlachos [53]. Fructose (4)
discharged one water molecule to form fructofuranosyloxocarbenium, then this oxocarbenium ion discharge H+ to form enolic compound (5) [51]. The enolic compound undergoes tautomerization
before elimination of two water molecules to form 5-HMF (7). 5HMF was then rehydrated with the presence of water in the system. A thorough mechanism of 5-HMF rehydration was previously
enlightened [54]. The nal products of the rehydration reaction of
5-HMF are LA (8) and formic acid (9). Furfural (10) might form in
the product mixture by the lost of formaldehyde from 5-HMF compound. The presence of LA, 5-HMF, formic acid, and furfural detected in the reaction product has supported the proposed reaction
mechanism.

100

Weight (wt%)

predicted LA yield. The LA yield at optimum conditions is 61.8%


indicating 1.3% error between the observed and predicted yield.
The error is considered small as the observed value is within the
5% signicance level.
In this study, LA is the main dehydration product from the glucose conversion process using 10% Fe/HY zeolite as catalyst. Yet,
other water soluble compounds were also detected such as 5HMF, formic acid, and furfural. At the optimum conditions, complete conversion of glucose was achieved, with 61.8%, 5.7%, 14.3%
and 10.4% of LA, 5-HMF, formic acid, and furfural yield, respectively. Other by-products could also be generated from the reaction,
such as lactic acid, as has been reported previously [23,49].
Besides, insoluble black solid residue regarded as humins was also
observed in all reactions.

H 2O

80
60

Cellulose
+ hemicellulose

40
20

Lignin
Ash

0
0

100

200

300

400

500

Temperature ( C)
Fig. 5. Thermal gravimetric analysis of OPF.

600

700

17

N.A.S. Ramli, N.A.S. Amin / Energy Conversion and Management 95 (2015) 1019
Table 4
LA production from various biomass feedstocks and catalysts.

a
b

Biomass

Cellulose content (%)

Catalyst

Water hyacinth
Wheat straw
Rice strawa
Giant reeda
Empty fruit bunch
Sorghum grain
Cotton strawb
Oil palm fronds
Oil palm frondsb
Empty fruit bunchb
Oil palm frondsa
Oil palm fronds
Oil palm fronds

26.3
40.4
46.1
36.6
41.1
73.8
42.6
45.2
45.2
38.2
45.2
45.2
45.2

H2SO4
H2SO4
S2O2
8 /ZrO2SiO2Sm2O3
HCl
Cr/HY zeolite
H2SO4
H2SO4
Fe/HY zeolite in [BMIM][Br]
Fe/HY zeolite in [BMIM][Br]
Fe/HY zeolite in [BMIM][Br]
Fe/HY zeolite
Fe/HY zeolite
Fe/HY zeolite

Reaction condition

LA yield (%)

Temperature (C)

Time (min)

Biomass

Theoretical

175
209
200
200
145
200
180
153
120
120
170
173
170

30
37
10
60
147
40
60
354
180
180
180
198
480

9.0
19.9
22.8
22.2
15.5
32.6
12.0
24.7
18.6
17.6
25.0
17.6
19.3

18.7
28.7
32.7
26.0
29.2
52.4
30.2
32.1
32.1
27.1
32.1
32.1
32.1

Efciency (%)

Ref

48.2
68.8
70.0
85.4
53.2
62.2
39.7
76.9
57.9
64.9
77.9
54.8
60.0

[20]
[5]
[7]
[16]
[34]
[14]
[55]
[40]
[39]
[39]
[56]
This study
This study

Pretreated biomass feedstock.


Biomass feedstocks were hydrolyzed to produce sugar before subjected to LA production.

3.4. Biomass conversion to LA


3.4.1. Determination of OPF composition
The determination of OPF cellulose content was conducted to
determine the highest theoretical LA yield and the efciency of
the process. Fig. 5 illustrates the thermal degradation curves of
OPF with three distinct stages of weight losses. The initial weigh
losses between 30 and 100 C is identied as evaporation of
residual water. The second stage at temperature range of 150
350 C, is volatilization of holocellulose. Meanwhile, the degradation of lignin is at 350520 C. The temperature ranges are aligned
with the previous TGA study on the determination of biomass
compositions [20,34]. The cellulose and hemicellulose contents in
OPF were determined using a two stage acid hydrolysis procedure
according to the NREL standard method. In the respective method,
cellulose and hemicellulose in biomass were degraded to glucose
and xylose subunits, respectively. From that basis, it was presumed
that the glucose and xylose compositions were corresponded to the
cellulose and hemicellulose in the OPF. From the analyses, OPF
contain 12.8% moisture, 11.2% lignin, 10.5% ash, 45.2% cellulose,
and 20.3% hemicellulose.
3.4.2. Comparative study
In this study, the potential of OPF to produce LA was inspected
over 10% Fe/HY zeolite catalyst. Since glucose is the model compound of cellulose, and OPF contains high percentage of cellulose,
the same optimum condition for glucose conversion to LA was
applied to OPF. The LA yield from OPF conversion at 173.4 C,
3.3 h, 0.93 g OPF, and 0.89 g 10% Fe/HY zeolite was 17.6%. The efciency of the process has been accounted as 54.8%. It is important
to note that higher efciency was achieved in this study, compared
to the previous study on oil palm biomass using chromium (Cr)
modied HY zeolite in aqueous system [34]. The conversion of
OPF at the optimum conditions obtained earlier reveals that
around half of the cellulose content has been converted into LA.
Basically, 100% efciency of biomass conversion to LA is not attainable. This is attributed to other possible parallel reactions that
could have occurred in the system including fragmentation and
polymerization. The biomass sample might decompose into other
compounds such as formic acid, lactic acid, humins, furfural, and
acetic acid [9]. Besides, biomass recalcitrance could also has stunted the hydrolysis process.
Table 4 lists the production of LA from different types of biomass feedstocks over several homogeneous and heterogeneous
catalysts. It is evident that the hydrolysis of biomass to LA yield
and process efciency over the Fe/HY zeolite in the present study

has been improved or comparable with other previous research.


At lower reaction temperature and lesser time, the process efciency of Cr/HY was comparable with Fe/HY zeolite, but the high price
and toxicity of Cr undermines the advantages of the system [34].
Higher LA yield is expected from biomass with higher cellulose
content [20]. The yields also relied on reaction conditions and catalyst acidity. Higher reaction temperature, prolonged reaction
time, and stronger acid catalyst could boost the LA yield.
Shorter reaction times were needed for reactions involving the
use of mineral acids such as H2SO4 and HCl [5,14,16,20,55]. However, applying mineral acid in biomass hydrolysis also raised drawbacks since the catalyst cannot be recycled for repeated uses and
the mineral acid can cause serious problems to the environment.
The initial hydrolysis of OPF to produce glucose using ionic liquid
and the utilization of ionic liquid instead of water as reaction medium have also been investigated [39,40]. The Fe/HY zeolite-ionic
liquid was effective as lower reaction temperature was sufcient
to increase the process efciency. Nevertheless, the use 10% Fe/
HY zeolite in aqueous medium is preferred from the economic
point of view. Besides, the 10% Fe/HY zeolite is a reusable catalyst
for LA production.
Reduction in cellulose crystallinity from biomass pretreatment
method could enhance the product yield [6,7]. Consequently, the
amorphous cellulose augmented the acid hydrolysis reaction as
the porous materials increase the accessibility of the catalyst to
cellulose. Pretreatment of biomass [7,16] and initial hydrolysis of
biomass to produce glucose [55], applied prior to LA production,
improved the process efciency within shorter reaction time. For
OPF conversion to LA over Fe/HY zeolite catalyst, the inuence of
pretreatment using combination of ionic liquid and aqueous acid
pretreatment has been evaluated [56]. The pretreated OPF has a
more porous and less crystalline structure, which is a desired feature for subsequent hydrolysis step for LA production. As a consequence, higher LA yield (25%) and process efciency (77.9%) were
attained. The inuence of other biomass pretreatment methods
for LA production should be considered in the future.
The optimum conditions of LA production from glucose conversion reaction have been employed for the hydrolysis of OPF. Due to
the complex structure of biomass, more severe condition would be
necessary to attain higher LA yield from OPF. Thus, additional testings have been conducted to evaluate the effect of prolonged reaction time on the LA yield from OPF conversion over 10% Fe/HY
zeolite. The reactions were carried out at 170 C, 1 g of OPF, and
1 g of 10% Fe/HY zeolite for reaction time from 2 to 10 h (Fig. S4).
The LA yield increased linearly from 8% to 19% for 2 to 8 h reaction
time. When the reaction was further conducted up to 10 h, a lower

18

N.A.S. Ramli, N.A.S. Amin / Energy Conversion and Management 95 (2015) 1019

Glucose

LA yield (%)

60

OPF

decrease in the LA yield with increasing cycle was mainly attributed to the decrease in the catalyst amount after subsequent
washings.

40

4. Conclusions
20

0
1

Run
Fig. 6. Reusability of 10% Fe/HY zeolite for LA production from glucose and OPF
conversions.

LA yield of 15% was obtained. The reduction in the LA yield was


possibly due to the increased formation of undesired by-products
such as humins.
3.5. Reusability of 10% Fe/HY zeolite
The reusability of heterogeneous catalyst is one of the main
advantages over homogeneous catalyst. In this study, the reusability of 10% Fe/HY zeolite has been tested for LA production from glucose and OPF conversions at 180 C for 3 h, using 1 g of feedstock,
and 1 g of 10% Fe/HY zeolite (for the rst cycle). After each catalytic
cycle, the catalyst was recovered by centrifugation and washed
with water. The catalyst was dried overnight at 120 C, calcined
at 400 C for 5 h to remove the adsorbed by-products such as
humins, and returned to the subsequent cycles. For the subsequent
cycles, the reaction were conducted at 180 C for 3 h, using 1 g of
feedstock, and the remaining catalyst from the previous cycle.
The 10% Fe/HY zeolite could be reused up to ve times with a
slight decrease of the LA yield (Fig. 6). The lower yield from the
reused catalyst was probably due to lower catalyst mass as a consequence of the multiple ltration and washing steps. As the
amount of catalyst decreased during the ltration steps, the number of active sites available for the reaction also reduced. Besides,
the presence of remaining OPF ash in the solid residue could
reduce the catalyst activity in the subsequent reaction runs. The
possible leaching of Fe from the 10% Fe/HY zeolite into the solution
was also studied. Sample of experiments were subjected to atomic
absorption spectroscopy (AAS). The amount of Fe ions was found to
be less than 1% of the starting Fe manifesting that the hydrolysis
reaction was mostly due to presence of Fe on the catalyst surface
rather than trace amount of leached Fe ions.
The regenerated 10% Fe/HY zeolite from glucose conversion
reaction has been examined using XRD, FTIR, FESEM, N2 physisorption, and NH3-TPD, and compared with the fresh 10% Fe/HY zeolite.
From the XRD proles, no signicant changes in the diffraction
peaks were noticed after the regeneration, which signied the
undisrupted 10% Fe/HY zeolite structure (Fig. S5). The FTIR spectra
demonstrated that all the functional groups remain unchanged
after the reaction (Fig. S6). The FESEM images (Fig. S7) indicate that
the external surface of the catalyst did not exhibit a signicant
change in texture after the reaction. The reduction in the surface
area for the regenerated 10% Fe/HY zeolite (538 m2/g) was only
2%. Meanwhile, the acidity of the spent 10% Fe/HY zeolite has been
measured using NH3-TPD. The reduction in acidity of the regenerated catalyst (2.63 mmol/g) was not signicant compared to the
fresh one (2.68 mmol/g). The number of acid sites of the regenerated 10% Fe/HY zeolite (4.88 lmol/m2) was also the same as the
fresh. The catalyst recovery, calculated by dividing the weight of
regenerated catalyst with the weight of fresh catalyst, was 88%
after the fth run. From the ndings, it can be concluded that the

The 10% Fe/HY zeolite has been demonstrated as an effective


solid catalyst for the production of LA. RSM coupled with Box
Behnken design was used for the optimization of LA yield from
glucose conversion whilst four parameters were evaluated for
LA production. The reaction parameters including reaction temperature, reaction time, glucose loading, and 10% Fe/HY zeolite
loading. The predicted values were in close agreement with the
experimental values which gave optimum conditions at 173.4 C,
3.3 h, 0.93 g glucose, and 0.89 g 10% Fe/HY zeolite loading with
61.8% LA yield. The conversion of OPF conducted at the same optimum conditions produced 17.6% LA yield, with 54.8% process efciency. From the leaching and reusability test, 10% Fe/HY zeolite is
an environmentally friendly and reusable solid acid catalyst for the
production of LA from glucose and OPF in aqueous medium.
Acknowledgements
The authors would like to extend their gratitude to Universiti
Teknologi Malaysia for the nancial support under the Research
University Grant (RUG) vote number 07H14. One of the authors,
(NASR) would like to thank the Ministry of Education (MOE) for
the fellowship under MyBrain15.
Appendix A. Supplementary material
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.enconman.2015.
02.013.
References
[1] Hegner J, Pereira KC, DeBoef B, Lucht BL. Conversion of cellulose to glucose and
levulinic acid via solid-supported acid catalysis. Tetrahedron Lett 2010;51:
2356.
[2] Werpy T, Petersen G. Top value added chemicals from biomass. Results of
screening for potential candidates from sugars and synthesis gas. Other
information: PBD, Medium: ED; Size, vol. I; 2004. 76pp [01.08.04].
[3] Galletti AMR, Antonetti C, De Luise V, Martinelli M. A sustainable process for
the production of [gamma]-valerolactone by hydrogenation of biomassderived levulinic acid. Green Chem 2012;14:688.
[4] Rackemann DW, Doherty WOS. The conversion of lignocellulosics to levulinic
acid. Biofuels, Bioprod Bioren 2011;5:198.
[5] Chang C, Cen P, Ma X. Levulinic acid production from wheat straw. Bioresour
Technol 2007;98:1448.
[6] Bevilaqua DB, Rambo MKD, Rizzetti TM, Cardoso AL, Martins AF. Cleaner
production: levulinic acid from rice husks. J Cleaner Prod 2013;47:96.
[7] Chen H, Yu B, Jin S. Production of levulinic acid from steam exploded rice straw
via solid superacid. Bioresour Technol 2011;102:3568.
[8] Galletti AMR, Antonetti C, Luise VD, Licursi D, Nasso NNoD. Levulinic acid
production from waste biomass. BioResources 2012;7:1824.
[9] Zeng W, Cheng D-G, Zhang H, Chen F, Zhan X. Dehydration of glucose to
levulinic acid over MFI-type zeolite in subcritical water at moderate
conditions. Reac Kinet Mech Cat 2010;100:377.
[10] Mller M, Harnisch F, Schrder U. Microwave-assisted hydrothermal
degradation of fructose and glucose in subcritical water. Biomass Bioenergy
2012;39:389.
[11] Hu L, Sun Y, Lin L, Liu S. 12-Tungstophosphoric acid/boric acid as synergetic
catalysts for the conversion of glucose into 5-hydroxymethylfurfural in ionic
liquid. Biomass Bioenergy 2012;47:289.
[12] Fan C, Guan H, Zhang H, Wang J, Wang S, Wang X. Conversion of fructose and
glucose into 5-hydroxymethylfurfural catalyzed by a solid heteropolyacid salt.
Biomass Bioenergy 2011;35:2659.
[13] Shen Y, Xu Y, Sun J, Wang B, Xu F, Sun R. Efcient conversion of
monosaccharides into 5-hydroxymethylfurfural and levulinic acid in InCl3
H2O medium. Catal Commun 2014;50:17.
[14] Fang Q, Hanna MA. Experimental studies for levulinic acid production from
whole kernel grain sorghum. Bioresour Technol 2002;81:187.

N.A.S. Ramli, N.A.S. Amin / Energy Conversion and Management 95 (2015) 1019
[15] Kang M, Kim SW, Kim J-W, Kim TH, Kim JS. Optimization of levulinic acid
production from Gelidium amansii. Renewable Energy 2013;54:173.
[16] Raspolli Galletti AM, Antonetti C, Ribechini E, Colombini MP, Nassi o Di Nasso
N, Bonari E. From giant reed to levulinic acid and gamma-valerolactone: a high
yield catalytic route to valeric biofuels. Appl Energy 2013;102:157.
[17] Kupiainen L, Ahola J, Tanskanen J. Kinetics of glucose decomposition in formic
acid. Chem Eng Res Des 2011;89:2706.
[18] Yan L, Yang N, Pang H, Liao B. Production of levulinic acid from bagasse and
paddy straw by liquefaction in the presence of hydrochloride acid. CLEAN
Soil, Air, Water 2008;36:158.
[19] Hayes DJ, Fitzpatrick S, Hayes MHB, Ross JRH. The bione process production
of levulinic acid, furfural, and formic acid from lignocellulosic feedstocks.
Bioreneries-industrial processes and products. Wiley-VCH Verlag GmbH;
2008. p. 139.
[20] Girisuta B, Danon B, Manurung R, Janssen LPBM, Heeres HJ. Experimental and
kinetic modelling studies on the acid-catalysed hydrolysis of the water
hyacinth plant to levulinic acid. Bioresour Technol 2008;99:8367.
[21] Li H, Govind KS, Kotni R, Shunmugavel S, Riisager A, Yang S. Direct catalytic
transformation of carbohydrates into 5-ethoxymethylfurfural with acidbase
bifunctional hybrid nanospheres. Energy Convers Manage 2014;88:1245.
[22] Peng L, Lin L, Zhang J, Zhuang J, Zhang B, Gong Y. Catalytic conversion of
cellulose to levulinic acid by metal chlorides. Molecules 2010;15:5258.
[23] Weingarten R, Kim YT, Tompsett GA, Fernndez A, Han KS, Hagaman EW, et al.
Conversion of glucose into levulinic acid with solid metal(IV) phosphate
catalysts. J Catal 2013;304:123.
[24] Jae J, Tompsett GA, Foster AJ, Hammond KD, Auerbach SM, Lobo RF, et al.
Investigation into the shape selectivity of zeolite catalysts for biomass
conversion. J Catal 2011;279:257.
[25] Xavier NM, Lucas SD, Rauter AP. Zeolites as efcient catalysts for key
transformations in carbohydrate chemistry. J Mol Catal A: Chem 2009;305:84.
[26] Li Z, Xie K, Slade RCT. Studies of the interaction between CuCl and HY zeolite
for preparing heterogeneous CuI catalyst. Appl Catal A 2001;209:107.
[27] Hu L, Sun Y, Lin L. Efcient conversion of glucose into 5-hydroxymethylfurfural
by Chromium(III) chloride in inexpensive ionic liquid. Ind Eng Chem Res
2011;51:1099.
[28] Tan M, Zhao L, Zhang Y. Production of 5-hydroxymethyl furfural from cellulose
in CrCl2/Zeolite/BMIMCl system. Biomass Bioenergy 2011;35:1367.
[29] Yuan Z, Xu C, Cheng S, Leitch M. Catalytic conversion of glucose to 5hydroxymethyl furfural using inexpensive co-catalysts and solvents.
Carbohydr Res 2011;346:2019.
[30] Wang S, Du Y, Zhang W, Cheng X, Wang J. Catalytic conversion of cellulose into
5-hydroxymethylfurfural over chromium trichloride in ionic liquid. Korean J
Chem Eng 2014;31:1786.
[31] Wang S, Du Y, Zhang P, Cheng X, Qu Y. One-pot synthesis of 5hydroxymethylfurfural directly from cottonseed hull biomass using
chromium (III) chloride in ionic liquid. Korean J Chem Eng 2014;31:2286.
[32] Mao L, Zhang L, Gao N, Li A. Seawater-based furfural production via corncob
hydrolysis catalyzed by FeCl3 in acetic acid steam. Green Chem 2013;15:727.
[33] Cornell JA. How to apply response surface methodology. American Society for
Quality Control; 1990.
[34] Yaaini N, Amin NAS, Asmadi M. Optimization of levulinic acid from
lignocellulosic biomass using a new hybrid catalyst. Bioresour Technol
2012;116:58.
[35] Yu XC, Sun DL, Li XS. Preparation of levulinic acid using cellulase and solid acid
synergistic hydrolysis for rice straw. Asian J Chem 2010;22:7113.
[36] Zhou C, Yu X, Ma H, He R, Vittayapadung S. Optimization on the conversion of
bamboo shoot shell to levulinic acid with environmentally benign acidic ionic
liquid and response surface analysis. Chin J Chem Eng 2013;21:544.

19

[37] Yaaini N, Amin NAS, Endud S. Characterization and performance of hybrid


catalysts for levulinic acid production from glucose. Microporous Mesoporous
Mater 2013;171:14.
[38] Ramli NAS, Amin NAS. Fe/HY zeolite as an effective catalyst for levulinic acid
production from glucose: characterization and catalytic performance. Appl
Catal B 2015;163:487.
[39] Ramli NAS, Amin NAS. Catalytic hydrolysis of cellulose and oil palm biomass in
ionic liquid to reducing sugar for levulinic acid production. Fuel Process
Technol 2014;128:490.
[40] Ramli NAS, Amin NAS. Catalytic conversion of oil palm fronds to levulinic acid
in ionic liquid. Appl. Mech. Mater. 2014;625:361.
[41] Carrier M, Loppinet-Serani A, Denux D, Lasnier J-M, Ham-Pichavant F, Cansell
F, et al. Thermogravimetric analysis as a new method to determine the
lignocellulosic composition of biomass. Biomass Bioenergy 2011;35:298.
[42] Laboratory NRE. Chemical analysis and testing standard procedure. Golden Co,
National Renewable Energy Laboratory (NREL), No. 002-005; 1996.
[43] Weingarten R, Tompsett GA, Conner Jr WC, Huber GW. Design of solid acid
catalysts for aqueous-phase dehydration of carbohydrates: the role of Lewis
and Brnsted acid sites. J Catal 2011;279:174.
[44] Jimnez-Morales I, Teckchandani-Ortiz A, Santamara-Gonzlez J, MairelesTorres P, Jimnez-Lpez A. Selective dehydration of glucose to 5hydroxymethylfurfural on acidic mesoporous tantalum phosphate. Appl
Catal B 2014;144:22.
[45] Jimnez-Morales I, Moreno-Recio M, Santamara-Gonzlez J, Maireles-Torres
P, Jimnez-Lpez A. Mesoporous tantalum oxide as catalyst for dehydration of
glucose to 5-hydroxymethylfurfural. Appl Catal B 2014;154155:190.
[46] Wan Omar WNN WNN, Saidina Amin NA. Optimization of heterogeneous
biodiesel production from waste cooking palm oil via response surface
methodology. Biomass Bioenergy 2011;35:1329.
[47] Cha JY, Hanna MA. Levulinic acid production based on extrusion and
pressurized batch reaction. Ind Crops Prod 2002;16:109.
[48] Li Y, Liu H, Song C, Gu X, Li H, Zhu W, et al. The dehydration of fructose to 5hydroxymethylfurfural efciently catalyzed by acidic ion-exchange resin in
ionic liquid. Bioresour Technol 2013;133:347.
[49] Chambon F, Rataboul F, Pinel C, Cabiac A, Guillon E, Essayem N. Cellulose
hydrothermal conversion promoted by heterogeneous Brnsted and Lewis
acids: Remarkable efciency of solid Lewis acids to produce lactic acid. Appl
Catal B 2011;105:171.
[50] Zhao H, Holladay JE, Brown H, Zhang C. Metal chlorides in ionic liquid solvents
convert sugars to 5-hydroxymethylfurfural. Science 2007;316:1597.
[51] Utami SP, Amin NS. Optimization of glucose conversion to 5hydroxymethylfulfural using [BMIM]Cl with ytterbium triate. Ind Crops
Prod 2013;41:64.
[52] Romn-Leshkov Y, Moliner M, Labinger JA, Davis ME. Mechanism of glucose
isomerization using a solid lewis acid catalyst in water. Angew Chem Int Ed
2010;49:8954.
[53] Caratzoulas S, Vlachos DG. Converting fructose to 5-hydroxymethylfurfural: a
quantum mechanics/molecular mechanics study of the mechanism and
energetics. Carbohydr Res 2011;346:664.
[54] Horvat J, Klaic B, Metelko B, unjic V. Mechanism of levulinic acid formation.
Tetrahedron Lett 1985;26:2111.
[55] Yang Z, Kang H, Guo Y, Zhuang G, Bai Z, Zhang H, et al. Dilute-acid conversion
of cotton straw to sugars and levulinic acid via 2-stage hydrolysis. Ind Crops
Prod 2013;46:205.
[56] Ramli NAS, Amin NAS. Optimization of oil palm fronds pretreatment using
ionic liquid for levulinic acid production. Jurnal Teknologi 2014;71:33.

You might also like