You are on page 1of 10

Hydrolysis of sugar cane bagasse using nitric acid:

a kinetic assessment
Antonio Rodrguez-Chong
a
, Josee Alberto Ramrez
b
, Gil Garrote
c
, Manuel Vaazquez
bd,
*
a
Department of Food Engineering, U. A. M. Mante, Universidad Aut oonoma de Tamaulipas, Boulevard Enrique C aardenas Gonz aalez 1201,
Col. Jardn, Cd. Mante, Tamaulipas 89840, Mexico
b
Department of Food Science and Technology, U. A. M. Reynosa-Aztl aan, Universidad Aut oonoma de Tamaulipas,
Apdo. Postal 1015, Reynosa, Tamaulipas 88700, Mexico
c
Departamento de Enxe~ nnera Qumica, Facultade de Ciencias, Universidade de Vigo Campus de Ourense,
Edicio Polit eecnico, As Lagoas, 32004 Ourense, Spain
d
Area de Tecnologa de los Alimentos, Departamento Qumica Analtica. Escuela Polit eecnica Superior,
Universidad de Santiago de Compostela, Campus de Lugo, 27002 Lugo, Spain
Received 18 November 2002; accepted 25 February 2003
Abstract
Sugar cane bagasse was hydrolysed using nitric acid at variable concentration (26%), reaction time (0300 min) and temperature
(100128 C). The concentration of sugars released (xylose, glucose and arabinose) and degradation products (acetic acid and
furfural) were determined and the kinetic parameters of mathematical models for predicting them in the hydrolysates were obtained.
The inuence of temperature was also studied using the Arrhenius equation. Applying the kinetic models obtained, the optimal
conditions selected were: 122 C, 6% HNO
3
and 9.3 min. Using these conditions, 18.6 g xylose/l; 2.04 g arabinose/l; 2.87 g glucose/l;
0.9 g acetic acid/l and 1.32 g furfural/l were obtained. Comparison of these results with those obtained using sulphuric and hy-
drochloric acids demonstrated that the nitric acid was the most ecient catalyst for hydrolysis.
2003 Elsevier Ltd. All rights reserved.
Keywords: Sugar cane; Bagasse; Xylose; Glucose; Arabinose; Kinetic; Nitric acid; Modeling; Acid hydrolysis
1. Introduction
Xylose is a sugar with application as a carbon source
in fermentation processes. The main application of
xylose is its bioconversion to xylitol, a functional
sweetener with important technological properties like
anticarcinogenicity, low caloric value and negative heat
of dissolution (Paraj oo, Domnguez, & Domnguez,
1998). The economic interest in xylitol production can
be enhanced if the solutions can be obtained from the
hydrolysis of low-cost lignocellulosic wastes. Among
these, sugar cane bagasse is a potential xylose source
whose hydrolysis to obtain xylose solutions has a double
consequence, the elimination of such a waste and the
generation of a value-added product thus increasing the
economy of the process. Other applications of this by-
product are as source of animal feed, energy, pulp,
paper and boards (Banerjee & Pandey, 2002).
Dilute acids lead to a limited hydrolysis called pre-
hydrolysis which consists in the hydrolysis of the
hemicellulosic fraction, remaining the cellulose and lig-
nin fractions almost unaltered. Sulphuric, hydrochloric,
hydrouoric or acetic acids are commonly employed as
catalysts. The product of the hydrolysis is a solution
containing mainly sugars such as xylose, glucose and
arabinose. Other compounds such as oligomers, furfural
and acetic acid are also released. Bonds in hemicellulosic
fraction are weaker than in cellulosic fraction. There-
fore, using selected operational conditions, it is possible
to hydrolyse almost quantitatively the hemicelluloses
leaving the cellulose and lignin in the solid residue,
which can be processed for the production of lactic acid
or ethanol (David, Fornasier, Greindl-Fallon, & Van-
lautem, 1985; Grethlein & Converse, 1991; Teixeira,
Linden, & Schroeder, 1999) through saccharication-
fermentation or for the production of paper pulp (Car-
aschi, Campana, & Curvelo, 1996). Nitric acid has been
Journal of Food Engineering 61 (2004) 143152
www.elsevier.com/locate/jfoodeng
*
Corresponding author. Tel.: +34-982-285900x22420; fax: +34-982-
241835.
E-mail address: vazquezm@lugo.usc.es (M. Vaazquez).
0260-8774/$ - see front matter 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/S0260-8774(03)00080-3
only applied up to now to the hydrolysis of hybrid
poplar (Luo-Caidian, Brink-David, & Blanch-Harvey,
2002) and corn starch (Yahiro, Shibata, Jia, Park, &
Okabe, 1997), but never in the hydrolysis of sugar cane
bagasse.
Xylose obtained from hemicellulosic hydrolysis of
sugar cane bagasse can be used for its chemical or bio-
technological conversion to xylitol (Rodrgues, Silva,
Prata, & Felipe, 1998), single cell protein (Nigam, 1998)
and ethanol (Gong, Chen, & Chen, 1993).
This work deals with the hydrolysis of sugar cane
bagasse using nitric acid for obtaining fermentable xy-
lose solutions. Kinetic models were developed to explain
the time course of xylose, glucose, arabinose and acetic
acid generated. The hydrolysis was optimised to obtain
xylose solutions with low concentration of growth in-
hibitors.
2. Materials and Methods
2.1. Raw material
The raw material used in experiments was sugar cane
bagasse collected in a local industry (Ingenio Azucarero
de Mante, Tamaulipas, Meexico). The sugar cane bagasse
was air dried, milled, screened to select the fraction of
particles with a size lower than 0.5 mm, homogenized in
a single lot and stored until needed.
2.2. Analysis of samples
Analyses of the main fractions (cellulose, hemicellu-
loses and Klason lignin) were carried out after a quan-
titative acid hydrolysis under standard conditions. The
solid residue of this hydrolysis is called Klason lignin
(Garrote, Domnguez, & Paraj oo, 1999a).
Treatments were performed in the range 100128 C
in media containing 2, 4 or 6 g HNO
3
/100 g liquor. All
experiments were performed using a liquor/solid ratio
(LSR) of 10 g liquor/g sugar cane bagasse on dry basis.
Samples were collected at several reaction times in the
range 0300 min. The experiments were performed in
nine experimental sets whose conditions are shown in
Table 1.
Samples of liquors were taken from the reaction
media and centrifuged. The pellets were washed twice
with sterile water and the supernatant used to determine
the composition of the hydrolysates.
The supernatant was diluted with water and analysed
by HPLC for glucose, xylose, arabinose and acetic acid.
The HPLC analyses were carried out using a HP-1100
(Agilent, Palo Alto, Ca) with a Transgenomic ION-300
column (oven temperature 45 C) with isocratic elu-
tion (ow rate 0.4 ml/min; mobile phase: H
2
SO
4
0.005
N) and a refraction index detector. Furfural was ana-
lysed by UVVis spectroscopy at 280 nm.
2.3. Statistical analysis
All experiments were carried out in triplicate and data
were expressed as average values. Non-linear regression
analyses of experimental data were performed with a
commercial optimization routine dealing with the
Newtonaas method (Solver, Microsoft Excel 2000, Mi-
crosoft Corporation, Redmond, WA) by minimizing the
sum of the squares of deviations between experimental
and calculated data, as reported previously (Garrote,
Domnguez, & Paraj oo, 2001a).
3. Results and discussion
The composition obtained for the sugar cane bagasse
was (weight percent on dry basis): Glucan, 38.9%;
Xylan, 20.6%; Araban, 5.56%; Klason lignin, 23.9%;
others, 11.0% (average values of three replicates,
error lower than 1% in all compounds). These values are
in the range found for this kind of material.
Table 1
Operating conditions employed for sugar cane bagasse hydrolysis with nitric acid and results of the tting for xylose concentration in the hydrolysis
of sugar cane bagasse with hydrochloric acid
Set Operating conditions Fitting the Saeman model Fitting the two-fraction model
Temperature
(C)
[HNO
3
]
(%)
Time (min) k
1
(min
1
) k
2
10
3
(min
1
)
r
2
a (g/g) k
1
(min
1
) k
2
10
3
(min
1
)
r
2
1 100 2 0300 0.006 3.9 0.952 0.40 0.029 0.0 0.940
2 100 4 0300 0.021 2.6 0.770 0.57 0.098 0.0 0.973
3 100 6 0300 0.035 3.0 0.809 0.65 0.170 0.6 0.995
4 122 2 0300 0.033 1.7 0.902 0.77 0.102 0.4 0.964
5 122 4 0300 0.060 1.3 0.972 0.83 0.529 0.0 0.984
6 122 6 0300 0.074 3.1 0.937 0.80 0.695 1.1 0.998
7 128 2 0300 0.033 0.6 0.927 0.85 0.260 0.0 0.900
8 128 4 0300 0.098 0.9 0.964 0.84 0.978 0.0 0.999
9 128 6 0300 0.053 1.5 0.817 0.81 1.47 0.4 0.998
144 A. Rodrguez-Chong et al. / Journal of Food Engineering 61 (2004) 143152
3.1. Composition of hydrolysates
The highest xylose concentration (21.0 g/l) was
reached after 180 min during the experiment of set 7. In
experiments performed with 6% HNO
3
, it was observed
that xylose concentration reached a maximum value and
then decreased with reaction time, thus conrming the
occurrence of decomposition reactions, probably lead-
ing to furfural (Fengel & Wegener, 1984). On the con-
trary, in experiments with lower concentration of nitric
acid, nal decrease in xylose concentration was not de-
tected.
Glucose concentration showed an initial fast increase
up to reach a nearly constant value (35 g/l) depending
on the operating conditions. In particular, it slightly
increased with temperature. No formation of hydroxy-
methylfurfural, a by-product able to inhibit the growth
of microorganism more than furfural, was detected
(Larsson et al., 1999, Pessoa, Mancilha, & Sato, 1996).
Arabinose, whose concentration exceeded 2 g/l, was
released more quickly than glucose and xylose. No
consumption of this sugar was observed under severe
conditions thus indicating that no degradation to fur-
fural took place.
The concentration of acetic acid, which is generated
for the hydrolysis of the acetyl groups of the hemicel-
luloses (Maloney, Chapman, & Baker, 1985), showed a
fast increase at the start of hydrolysis. Besides, it in-
creased slightly with temperature and hardly with nitric
acid concentration. The maximum concentrations va-
ried between 2 g/l (experiment 1) and 4 g/l (experiment
9). Acetic acid can be an inhibitor of microbial growth
when present from 410 g/l (Ferrari, Neirotti, Albornoz,
& Saucedo, 1992; Lawford & Rousseau, 1998) because it
enters the cell membrane and decreases intracellular pH,
thus aecting the metabolism of the microorganism
(Maiorella, Blanch, & Wilke, 1983; van Zyl, Prior, & Du
Preez, 1991). However, it has been reported that acetic
acid concentration of 910 g/l can stimulate the growth
of microorganism (Palmqvist, Almeida, & Hahn-
Haagerdal, 1999). In our study, the maximum acetic acid
concentrations (AcH) were close to the low limit of the
toxic eect.
3.2. Kinetic models
The use of dilute acids in the selected range of tem-
perature mainly conducts to the hydrolysis of polysac-
charides. The analytical methods applied can determine
sugar concentrations (glucose, xylose and arabinose)
and these are expressed as homopolymers in the original
lignocellulosic material (glucan, xylan, araban and ace-
tyl groups). Glucan (mainly cellulose), xylan and araban
are linked in the hemicelluloses, forming the araban-
oxylan, the main heteropolymer of agricultural mate-
rials (Garrote, Domnguez, & Paraj oo, 1999b).
A stringent kinetic study of the acid hydrolysis of
lignocellulosic materials is very dicult due to several
factors: (i) protection against the attacks of chemicals
(or biologicals) to the structure of the whole cells, (ii)
dicult access of protons to the raw material caused by
lignin hydrophobicity, (iii) interaction with other com-
ponents, (iv) presence of strong bonds in the raw ma-
terial (between units of xylose, acetyl groups, uronic
acids, other sugars, lignin, etc), (v) variable exposition of
hemicelluloses surface to the chemical attack along the
reaction (Abatzoglou, Koeberle, Chornet, Overend, &
Koukios, 1990; Carrasco & Roy, 1992; Conner, 1984).
Satisfactory results were obtained in kinetic studies
applying simplications, such as the use of principles
and laws valid for homogeneous systems, or the as-
sumption that the main fractions of the lignocellulosic
materials do not react among them. Therefore, the use
of pseudohomogeneous kinetic models in liquid phase
with rst order reactions were extended (Maloney et al.,
1985). These models began with the work of Saeman for
the hydrolysis of douglas r wood using sulphuric acid
(Saeman, 1945). In this research the hydrolysis of cel-
lulose was studied and the following model was demon-
strated to apply:
Cellulose !Glucose !Decomposition products 1
where both are rst order reactions. The model of Sae-
man was also applied to the hydrolysis of the hemicellu-
losic fraction (Grant, Han, Anderson, & Frey, 1977;
Teellez-Luis, Ramrez, & Vaazquez, 2002). Generalizing,
the model of Saeman can be applied to other homo-
polymers:
Polymers !
k
1
Monomers !
k
2
Decomposition products
2
where k
1
and k
2
are the kinetic coecient of the reactions
of monomer release and decomposition, respectively,
both having units of reciprocal time. Decomposition
products can be furfural, hydroxymethylfurfural, formic
acid, levulinic acid, etc. Solving the dierential equations
for an isothermal reaction, the following model predicts
the concentration of monomers:
M M
0
e
k
2
t
P
0
k
1
k
2
k
1
e
k
1
t
e
k
2
t
3
where M and P are the concentrations of monomer and
polymer expressed in g/l, t is time and subscript 0 indi-
cates initial conditions. In this work, Eq. (3) has been
applied to model the hydrolysis of sugar cane bagasse
with nitric acid.
3.3. Kinetic modelling of xylose concentration
Xylose is the main product of the hydrolysis of sugar
cane bagasse. To apply the Saeman model, the value of
M
0
was 0 g/l while P
0
was determined, assuming a total
A. Rodrguez-Chong et al. / Journal of Food Engineering 61 (2004) 143152 145
conversion of xylan to xylose in the raw material with-
out degradation, by Eq. (4):
P
0

150
132
CXn
0
=100
WSR
1000 23:4 g xylose=l 4
where CXn
0
is the initial composition for xylan (20.6 g
xylan/100 g sugar cane bagasse on dry basis), WSR is
the water/solid ratio (10 g water/g sugar cane bagasse)
and 150/132 is the ratio of the stoichiometric coe-
cients.
Posterior research to that of Saeman showed that
better ts to the kinetic models could be obtained con-
sidering two fractions inside the polymer (Kim & Lee,
1987), the one easily susceptible to the hydrolysis (fast
fraction) and the other dicultly susceptible (slow
fraction). The parameter a, dened as the mass ratio of
fast xylan to total xylan, usually takes values varying in
the range 0.51 g/g. The main reasons for using two
fractions are: (i) transport limitations of mass and
energy in the reactive mass, which are more important
when bigger are the particle size and lower the WSR; (ii)
presence of dierent fractions of hemicelluloses with
dierent intrinsic reactivity due to dierences in struc-
ture, accessibility, etc; (iii) presence of uronic acid with
lower reactivity than xylose that decreases the rate of
hemicelluloses solubilization, (iv) dierent reaction rate
of the fraction linked with lignin; (v) variations in the
surface of the interphase water-hemicelluloses along the
reaction (Abatzoglou, Chornet, Belkacemi, & Overend,
1992; Carrasco, Chornet, Overend, & Heitz, 1987;
Carrasco & Roy, 1992; Conner, 1984; Maloney et al.,
1985). Eq. (3) was modied according with the model of
two fractions to yield:
M M
0
e
k
2
t
aP
0
k
1
k
2
k
1
e
k
1
t
e
k
2
t
5
The results of tting the kinetic parameters of both the
Saeman model and the Two-fraction model are shown
In Table 1. It can be concluded, comparing the values of
parameter r
2
, that the model of two-fractions ts better.
Fig. 1 shows the comparison between experimental data
of xylose concentration and those calculated with the
two-fraction model.
It is note worthy that k
1
increased with temperature
and the acid concentration, while the values of k
2
were
zero or close to zero, suggesting that the rate of xylose
release was high and the degradation reaction negligible
or very slow.
The value of parameter a was in the range 0.400.85
g/g, with an average of 0.72 g/g, which is in accordance
with those reported for acid hydrolysis using other
lignocellulosic materials. For example, values of a in the
range 0.580.80 g/g were reported for the hydrolysis of
oak (Kim & Lee, 1987), 0.84 g/g for corn and 0.86 g/g
for sunower seeds (Eken-Sarac ogglu, Ferda, Dilmac , &
Cavus ogglu, 1998). It was often found that the parameter
a depends on the operational conditions. In the hydro-
lysis of Pinus pinaster at atmospheric pressure the values
of a were in the range 0.570.63 g/g while at autoclave
pressure were in the range 0.860.87 g/g (Garrote,
Domnguez, & Paraj oo, 2001b; Paraj oo, Santos, & del Ro,
1995b). In our work, the parameter a increased with
temperature, being the average value 0.54 g/g at 100 C,
0.80 g/g at 122 C y 0.83 g/g at 128 C.
The kinetic coecients can be correlated to obtain an
equation by which the inuence of the temperature on
the hydrolysis is taken into account. Traditionally, the
equation of Arrhenius is used for this purpose:
k
i
k
i0
e

Ea
RT
6
where k
i
is the kinetic coecient, k
i0
is a pre-exponential
factor, having the same units as k
i
, E
a
is the activation
energy (kJ/mol), R is the gas constant, 8.3143 10
3
kJ/
(mol K) and T is the absolute temperature (K).
0
4
8
12
16
20
24
0 100 200 300
0
4
8
12
16
20
24
0 100 200 300
0
4
8
12
16
20
0 100 200 300
2% HNO
3
4% HNO
3
6% HNO
3
2%
3
6% HNO
3
Fig. 1. Experimental and predicted dependence of xylose concentration
on time at dierent temperatures (

100 C;

122 C; r
128 C).
146 A. Rodrguez-Chong et al. / Journal of Food Engineering 61 (2004) 143152
Eq. (6) was applied to the kinetic coecients pre-
viously obtained. As frequently found in this kind of
hydrolysis (Garrote, Domnguez, & Paraj oo, 2001c;
Maloney et al., 1985), the values of k
2
were zero or too
small; therefore it was not possible to get a good tting
because they are greatly aected by experimental errors.
Table 2 shows the results of tting the kinetic coe-
cients k
1
to the equation of Arrhenius. The tting was
performed for each HNO
3
concentration, sets 1-4-7, 2-5-
8 and 3-6-9. The average values of E
a
and ln k
10
were 104
kJ/mol and 30.4 min
1
, respectively.
Sulphuric acid (Aguilar et al., 2002) and hydrochloric
acid (Bustos et al., 2003) have also been studied for the
hydrolysis of sugar cane bagasse. Table 2 shows that the
values of activation energy and pre-exponential factor
are similar for the three acid (the average values of E
a
and ln k
10
were 105 kJ/mol and 30.3 min
1
), being the
small dierences due to the experimental error. This
demonstrates that the kind of acid had no inuence on
the kinetic parameters of the hydrolysis of sugar cane
bagasse. These values are in the range found for other
lignocellulosics in the literature. In particular, for
models with two fractions for xylan, E
a
of the easy
fraction took values around 127 kJ/mol (Kim, Yum, &
Park, 2000), 120 kJ/mol (Aguilar et al., 2002) and 96.3
kJ/mol for hard woods (Eken-Sarac ogglu et al., 1998)
and 80.3 and 92.3 kJ/mol for agricultural wastes (Paraj oo,
Santos, & del Ro, 1995a).
3.4. Kinetic modelling of glucose concentration
During the acid hydrolysis of lignocellulosic mate-
rials, glucose is also released, although in lower con-
centrations than xylose, either from cellulose or
hemicellulosic heteropolymers. However, since the for-
mer polymer is very resistant to dilute acids, its hydro-
lysis can be neglected. Nevertheless, analytical methods
can not distinguish if glucose comes from cellulose or
hemicelluloses.
In a rst approach, the model of Saeman was applied
to model glucose concentration. The potential concen-
tration of glucose Gn
0
was determined as it was done
for xylose, resulting Gn
0
43:2 g/l. The results of tting
by the Saeman model (Table 3) were unsatisfactory,
mainly showing r
2
< 0:9, likely due to the assumption
that all cellulose would be susceptible of acid hydrolysis
under the operating conditions assayed. However, This
was unlikely because the cellulose forms a structure very
resistant in the lignocellulose materials and therefore
dicult of reacting. On the other hand, hemicelluloses
present an amorphous structure that contributes to easy
hydrolysis in acid medium (values of a close 1).
In a second approach, it was applied the model of
two-fractions (proposed for xylan in Eq. 5), which con-
siders that only a fraction of the polymer is susceptible
of hydrolysis. Therefore, a
G
was dened as the glucose
fraction susceptible of hydrolysis (g of hydrolysable
Table 2
Results of tting kinetic coecient k
1
(rate of xylose generation) for sugar cane bagasse hydrolysis with nitric acid to the equation of Arrhenius and
values from the literature with sulphuric and hydrochloric acids
Set of HNO
3
[HNO
3
] (%) ln k
10
(min
1
) E
a
(kJ/mol) r
2
1-4-7 2 26.1 85.6 0.941
2-5-8 4 29.9 100 0.995
3-6-9 6 35.3 125 0.976
Average 30.4 104
Other acids Reference
H
2
SO
4
31.6 109 Aguilar, Ramrez, Garrote, and Vaazquez (2002)
HCl 28.9 102 Bustos, Ramrez, Garrote, and Vaazquez (2003)
Table 3
Results of the tting for glucose concentration in the hydrolysis of sugar cane bagasse with nitric acid
Set Saeman model Two-fraction model
k
1
10
3
(min
1
) k
2
10
3
(min
1
) r
2
a
G
(g/g) k
1
(min
1
) k
2
10
3
(min
1
) r
2
1 7.6 16.4 0.993 0.04 0.021 0.0 0.996
2 1.2 15.8 0.899 0.06 0.026 0.0 0.942
3 1.6 17.7 0.943 0.07 0.031 0.0 0.984
4 0.5 0.4 0.990 0.53 0.001 0.0 0.990
5 1.4 5.6 0.859 0.16 0.011 0.0 0.871
6 3.8 22.9 0.801 0.10 0.127 0.5 0.997
7 0.9 12.7 0.871 0.21 0.004 6.3 0.888
8 2.0 14.9 0.761 0.09 0.042 0.1 0.994
9 2.0 12.6 0.780 0.10 0.034 0.0 0.976
A. Rodrguez-Chong et al. / Journal of Food Engineering 61 (2004) 143152 147
glucan/total glucan) in a similar equation, which gives
the concentration of glucose versus time.
G G
0
e
k
2
t
a
G
Gn
0
k
1
k
2
k
1
e
k
1
t
e
k
2
t
7
where G
0
0 g/l is the glucose concentration at time 0;
Gn
0
43:2 g/l, the glucose concentration corresponding
to hypothetical quantitative conversion of glucan to
glucose; k
1
is the kinetic coecient of hydrolysis of glu-
can to glucose and k
2
is the kinetic coecient of glucose
decomposition to hydroxymethylfurfural.
Table 3 and Fig. 2 show the results of tting glucose
concentration using this model for which we obtained r
2
values higher those observed with the one fraction
model. While k
1
increased with nitric acid concentration,
the decomposition of glucose was negligible, being the
coecient k
2
close to 0 in the most of experiments.
The coecient a
G
was not signicantly aected either
by temperature or nitric acid concentration, and showed
an average value of 0.15 g/g, corresponding to a sus-
ceptible glucan of 5.9 g /100 g sugar cane bagasse (on
dry basis), which means that the treatment with nitric
acid was selective toward the xylan hydrolysis.
3.5. Kinetic modelling of arabinose concentration
Arabinose is a sugar formed from arabinoxylans,
hemicellulosic heteropolymers found in agricultural
materials like sugar cane bagasse. In arabinoxylans, the
amount of xylose is higher than that of arabinose. The
same kinetic models developed for glucose were applied
(model of Saeman and two-fractions considering only
one of which being susceptible of hydrolysis). The pa-
rameter a
A
is the ratio between susceptible araban and
total araban.
Table 4 shows the results of tting both models.
While only in two sets tting using the Saeman model
showed statistical signicance r
2
> 0:9, the model of
two fractions tted better (all experiments showed
r
2
> 0:9). Fig. 3 shows the experimental and calculated
values using the two-fraction model for arabinose con-
centration.
Kinetic coecients for arabinose decomposition
(probably to furfural) were zero in the two-fraction
model for almost all experiments, while the value of a
A
varied from 0.31 to 0.42 g/g. The fraction of susceptible
araban was practically unaected by temperature, in-
0
1
2
3
4
0 100 200 300
0
2
4
6
8
10
0 100 200 300
0
1
2
3
4
5
0 100 200 300
2% HNO
3
4% HNO
3
6% HNO
3
0
6% HNO
3
Fig. 2. Experimental and predicted dependence of glucose concentra-
tion on time at dierent temperatures ( 100 C; d 122 C; r
128 C).
Table 4
Results of the tting for arabinose concentration in the hydrolysis of sugar cane bagasse with nitric acid
Set Saeman model Two-fraction model
k
1
10
3
(min
1
) k
2
10
3
(min
1
) r
2
a
A
(g/g) k
1
(min
1
) k
2
10
3
(min
1
) r
2
1 8.3 6.6 0.810 0.31 0.063 0.0 0.972
2 10.3 5.0 0.625 0.36 0.187 0.0 0.922
3 14.3 5.7 0.869 0.40 0.659 0.1 0.992
4 12.7 5.9 0.763 0.37 0.143 0.0 0.992
5 14.3 4.6 0.677 0.42 0.917 0.0 0.989
6 11.8 5.7 0.880 0.37 1.40 0.0 0.994
7 10.2 4.2 0.943 0.37 0.058 0.0 0.986
8 12.5 5.3 0.733 0.38 0.197 0.0 0.981
9 14.4 7.3 0.959 0.39 0.117 0.0 0.995
148 A. Rodrguez-Chong et al. / Journal of Food Engineering 61 (2004) 143152
creased slightly with the concentration of nitric acid, and
showed an average value of 0.38 g/g.
3.6. Kinetic modelling of acetic acid concentration
Some hemicellulosic monomers like xylose are linked
to acetyl groups (Ac), which can be hydrolysed to acetic
acid in acid media. Consistently with this knowledge,
our results showed that the acetic acid concentration
increased until a constant value, according to the simple
model (Garrote et al., 2001c; Lawford & Rousseau,
1998; Maloney et al., 1985):
Acetyl groups !
k
1
Acetic acid 8
The following dierential equation can be obtained on
the basis of this reaction model.
dAc
dt
k
1
Ac 9
Solving dierential equation leads to Eq. (10), which
gives the AcH as a function of time t.
AcH Ac
0
1 e
k
1
t
10
where Ac
0
is the potential concentration of acetyl
groups, expressed as acetic acid and introduced as a
regression parameter, and k
1
the rate of acetic acid
generation (min
1
).
Table 5 shows the coecients obtained by the tting.
Fig. 4 represents the experimental data and the predic-
tion of the model. Ac
0
increased with nitric acid con-
centration and temperature and varied in the range
1.835.87 g/l (average value of 3.35 g/l).
3.7. Kinetic modelling of furfural concentration
Furfural is generated as a decomposition product of
pentoses during the hydrolysis of sugar cane bagasse.
On the basis of the experimental data for furfural, a
model similar to that for acetic acid can be proposed
that expresses the furfural concentration F as a func-
tion of time t.
F F
0
1 e
k
1
t
11
where F
0
, introduced as a regression parameter, is the
potential concentration of furfural and k
1
the rate of
furfural generation (min
1
). Experimental data and
predicted values are compared in Fig. 5, while Table 6
lists the kinetic and statistical parameters obtained by
the tting of furfural generated during the hydrolysis.
Although F
0
increased either with temperature or
with nitric acid concentration, its values varied within a
narrow range (0.691.36 g/l) and appeared to be ade-
quate for posterior fermentation.
3.8. Overall Optimization
The main goal of this research was to develop kinetic
equations that allow determining optimal conditions for
the hydrolysis of sugar cane bagasse to obtain fer-
mentable sugar solutions. It is required that the sugar
solutions have high sugar concentration (carbon source
0
0.5
1
1.5
2
2.5
3
0 100 200 300
0
0.5
1
1.5
2
2.5
0 100 200 300
0
0.5
1
1.5
2
2.5
3
0 100 200 300
2% HNO
3
4% HNO
3
6% HNO
3
6% HNO
3
Fig. 3. Experimental and predicted dependence of arabinose concen-
tration on time at dierent temperatures (

100 C;

122 C; r
128 C).
Table 5
Results of the tting for acetic acid concentration in the hydrolysis of
sugar cane bagasse with nitric acid
Set Ac
0
(g/l) k (min
1
) r
2
1 1.83 0.071 0.973
2 2.39 0.618 0.714
3 3.00 1.39 0.567
4 3.07 0.031 0.849
5 2.61 0.062 0.949
6 5.87 0.018 0.985
7 3.62 0.032 0.910
8 3.64 0.137 0.990
9 4.16 0.082 0.991
A. Rodrguez-Chong et al. / Journal of Food Engineering 61 (2004) 143152 149
for micro-organism growth) and low concentration of
growth inhibitors (acetic acid and furfural). Using Eqs.
(5), (7), (10) and (11) and the values of kinetic para-
meters showed in Tables 25, operational conditions
were optimised. Table 7 shows the optimal time found
for each set of operation conditions (temperature and
acid concentration) as well as the composition of the
hydrolysates obtained under these conditions.
Hydrolysis with nitric acid allowed obtaining high
xylose concentrations (around 20 g/l) at 122 and 128 C,
while arabinose concentration varied between 1.04 g/l
(128 C, 6% acid and 5.6 min) and 2.33 g/l (122 C, 4%
acid and 5.8 min), glucose concentration between 0.79 g/
l (128 C, 6% acid and 6 min) and 2.87 g/l (122 C, 6%
acid and 9.3 min), and that of acetic acid between 0.91 g/
l (122 C, 6% acid and 9.3 min) and 3.00 g/l (100 C, 6%
acid and 33.0 min). Overall optimum conditions were
considered to be those (122 C, 6% nitric acid and 9.3
min) able to allow obtaining 23.5 g sugars/l (18.6 g xy-
lose/l, 2.04 g arabinose/l and 2.87 g glucose/l) and only
0.9 g acetic acid/l and 1.32 g furfural/l. Sugar solutions
with similar composition obtained from agricultural
materials (Garrote et al., 2001c) were successfully ap-
plied as carbon source in the fermentation to xylitol
(Rivas, Domnguez, Domnguez, & Paraj oo, 2002).
0
1
2
3
4
5
6
0 100 200 300
0
1
2
3
4
0 100 200 300
0
1
2
3
4
5
0 100 200 300
2% HNO
3
4% HNO
3
6% HNO
3
Fig. 4. Experimental and predicted dependence of acetic acid con-
centration on time at dierent temperatures (

100 C;

122 C;
r 128 C).
0
0.5
1
1.5
0 100 200 300
0
0.4
0.8
1.2
0 100 200 300
0
0.5
1
1.5
0 100 200 300
2% HNO
3
4% HNO
3
6% HNO
3
00
00
Fig. 5. Experimental and predicted dependence of furfural concen-
tration on time at dierent temperatures (

100 C;

122 C; r
128 C).
Table 6
Results of the tting for furfural concentration in the hydrolysis of
sugar cane bagasse with nitric acid
Set F
0
(g/l) k (min
1
) r
2
1 0.69 0.032 0.989
2 0.84 0.064 0.971
3 0.93 0.100 0.992
4 1.01 0.092 0.978
5 1.26 0.102 0.975
6 1.32 0.883 0.981
7 1.10 0.102 0.865
8 1.30 0.120 0.967
9 1.36 0.965 0.987
150 A. Rodrguez-Chong et al. / Journal of Food Engineering 61 (2004) 143152
3.9. Comparison of dilute acids
Several diluted acids have been compared as cata-
lyists for the hydrolysis of sugar cane bagasse. The e-
ciencies of the nitric acid determined in this work have
been compared with those reported for sulphuric acid
(Aguilar et al., 2002) and hydrochloric acid (Bustos
et al., 2003) under the same operating conditions. The
eciency has been dened as the ratio:
E
P
S
P
I
12
where
P
S is the sum of the concentrations of sugars in
the hydrolysates (xylose and glucose) and
P
I is the sum
of the concentrations of all inhibitors in the hydrolysates
(acetic acid and furfural). Arabinose was not considered
because it was not determined for all acids. Therefore,
on the basis of the values of E, a direct comparison has
been performed to determine the eect of the kind of
acid in the hydrolysis of sugar cane bagasse.
Table 8 shows the E values for HNO
3
(this work),
H
2
SO
4
(Aguilar et al., 2002) and HCl (Bustos et al.,
2003). Nitric acid needed lower time under optimal
conditions than sulphuric or hydrochloric acids. All
three acids gave solutions with high concentration of
sugars and low concentrations of growth inhibitors. The
values of E increased with temperature (average values
of 5.0, 5.1 and 7.5 at 100, 122 and 128 C, respectively)
and with acid concentration (average values of 5.2, 5.8
and 6.6 for 2%, 4% and 6% nitric acid, respectively).
The average values of E were 6.1, 5.3 and 6.2 for
nitric, hydrochloric and sulphuric acids, respectively.
HCl was the least ecient acid due to the large amount
of inhibitors generated, whereas H
2
SO
4
was little more
ecient than nitric acid. Selecting only the experiments
of industrial interest (with more than 20 g sugar/l), E
values of 7.4, 5.3 and 5.2 were obtained for nitric acid,
hydrochloric acid and sulphuric acid, respectively.
It can be concluded that nitric acid was the most ef-
cient in sugar cane bagasse hydrolysis (high E Value),
since the experiments with high E values for nitric acid
were those ensuring high sugar concentration and low
concentration of inhibitors.
References
Abatzoglou, N., Chornet, E., Belkacemi, K., & Overend, R. P. (1992).
Phenomenological kinetics of complex systems: The development
of a generalized severity parameter and its application to ligno-
cellulosics fractionation. Chemical Engineering Science, 47, 1109
1112.
Abatzoglou, N., Koeberle, P. G., Chornet, E., Overend, R. P., &
Koukios, E. G. (1990). An application to medium consistency
suspensions of hardwoods using a plug ow reactor. Canadian
Journal of Chemical Engineering, 68, 627638.
Table 7
Composition of the hydrolysates of sugar cane bagasse obtained under dierent conditions
Temp. (C) [HNO
3
] (weight %) Time (min) Xylose (g/l) Arabinose (g/l) Glucose (g/l) Acetic acid (g/l) Furfural (g/l)
100 2 300 9.35 1.75 1.70 1.83 0.698
4 300 13.3 2.03 2.59 2.39 0.838
6 33.0 14.8 2.20 1.81 3.00 0.892
122 2 54.0 17.6 2.07 1.21 2.49 1.00
4 15.8 19.5 2.33 1.09 1.63 1.01
6 9.3 18.6 2.04 2.87 0.91 1.32
128 2 32.1 19.8 1.76 1.00 2.33 1.05
4 9.8 19.7 1.81 1.30 2.69 0.901
6 5.6 18.8 1.04 0.79 1.54 1.36
Table 8
Values of E (catalytic eciency) for the experiments in obtained under dierent conditions
Temperature (C) [acid] (weight %) HNO
3
HCl
a
H
2
SO
4
a
Time (min) E (g/g) Time (min) E (g/g) Time (min) E (g/g)
100 2 300 4.37 300 5.56 167.5 4.69
4 300 4.92 300 5.52 300 4.91
6 33 4.27 300 5.06 300 5.42
122 2 54 5.39 300 2.79 24.1 5.90
4 15.8 7.80 300 2.07 25 5.27
6 9.3 9.63 107.1 2.73 18.1 4.55
128 2 32.1 6.15 51.1 5.14 20.1 6.94
4 9.8 5.85 22.2 8.76 11.4 7.48
6 5.6 6.76 7.6 10.1 5.6 10.5
a
Based on data from Aguilar et al. (2002) using H
2
SO
4
and from Bustos et al. (2003) using HCl.
A. Rodrguez-Chong et al. / Journal of Food Engineering 61 (2004) 143152 151
Aguilar, R., Ramrez, J. A., Garrote, G., & Vaazquez, M. (2002).
Kinetic study of the acid hydrolysis of sugar cane bagasse. Journal
of Food Engineering, 55, 309318.
Banerjee, R., & Pandey, A. (2002). Bio-industrial applications of
sugarcane bagasse: A technological perspective. International Sugar
Journal, 104(1238), 64, 6667.
Bustos, G., Ramrez, J. A., Garrote, G., & Vaazquez, M. (2003).
Modelling of the hydrolysis of sugar cane bagasse using hydro-
chloric acid. Applied Biochemistry and Biotechnology, 104(1),
5168.
Caraschi, J. C., Campana, S. P., & Curvelo, A. A. S. (1996).
Preparation and characterization of dissolving pulps obtained
from sugar cane bagasse. Polmeros: Ciencia e Tecnologia, 6, 2429.
Carrasco, F., Chornet, E., Overend, R. P., & Heitz, M. (1987).
Fractionnement de deux bois tropicaux (Eucalyptus et Wapa) par
traitment thermomecanique en phase aqueuse. Partie II: Charac-
teristiques chimiques des reesidus et considerations cinetiques sur la
solubilisation des hemicelluloses. Canadian Journal of Chemical
Engineering, 65, 7177.
Carrasco, F., & Roy, C. (1992). Kinetic study of dilute-acid
prehydrolysis of xylan-containing biomass. Wood Science Techno-
logy, 26, 189208.
Conner, A. H. (1984). Kinetic modelling of hardwood prehydrolysis.
Part I: Xylan removal by water prehydrolysis. Wood and Fiber
Science, 16, 268277.
David, C., Fornasier, R., Greindl-Fallon, C., & Vanlautem, N. (1985).
Enzymatic hydrolysis and bacterian hydrolysis-fermentation of
Eucalyptus wood pretreated with sodium hypochlorite. Biotech-
nology and Bioengineering, 26, 15911595.
Eken-Sarac ogglu, N., Ferda, D., Dilmac , G., & Cavus ogglu, H. (1998).
A comparative kinetic study of acidic hemicellulose hydrolysis in
corn cob and sunower seed hulls. Bioresource Technology, 65,
2933.
Fengel, D., & Wegener, G. (1984). Wood. Chemistry, ultrastructure,
reactions. Berlin: Ed. Walter de Gruyter.
Ferrari, M. D., Neirotti, E., Albornoz, C., & Saucedo, E. (1992).
Ethanol production from Eucalyptus wood hemicellulose hydro-
lysate by Pichia stipitis. Biotechnology and Bioengineering, 40,
753759.
Garrote, G., Domnguez, H., & Paraj oo, J. C. (1999a). Mild autohy-
drolysis: An environmentally friendly technology for xylooligosac-
charide production from wood. Journal of Chemical Technology
and Biotechnology, 74, 11011109.
Garrote, G., Domnguez, H., & Paraj oo, J. C. (1999b). Hydrothermal
processing of lignocellulosic materials. Holz als Roh und Werksto,
57, 191202.
Garrote, G., Domnguez, H., & Paraj oo, J. C. (2001a). Kinetic modeling
of corncob autohydrolysis. Process Biochemistry, 36, 571578.
Garrote, G., Domnguez, H., & Paraj oo, J. C. (2001b). Generation of
xylose solutions from Eucalyptus globulus wood by autohydroly-
sis-posthydrolysis processes: Posthydrolysis kinetics. Bioresource
Technology, 79, 155164.
Garrote, G., Domnguez, H., & Paraj oo, J. C. (2001c). Manufacture of
xylose-based fermentation media from corncobs by posthydrolysis
of autohydrolysis liquors. Applied Biochemistry and Biotechnology,
95, 195207.
Gong, C. S., Chen, C. S., & Chen, L. F. (1993). Pretreatment of sugar
cane bagasse hemicellulose hydrolyzate for ethanol production by
yeast. Applied Biochemistry and Biotechnology, 39/40, 8388.
Grant, G. A., Han, Y. W., Anderson, A. W., & Frey, K. L. (1977).
Kinetics of straw hydrolysis. Developments in Industrial Microbio-
logy, 18, 599611.
Grethlein, H. E., & Converse, A. O. (1991). Common aspects of acid
prehydrolysis and steam explosion for pretreating wood. Biore-
source Technology, 36, 7782.
Kim, S. B., & Lee, Y. Y. (1987). Kinetics in acid-catalyzed hydrolysis
of hardwood hemicellulose. Biotechnology Bioengineering Sympo-
sium, 17, 7184.
Kim, S. B., Yum, D. M., & Park, S. C. (2000). Step-change variation of
acid concentration in a percolation reactor for hydrolysis of
hardwood hemicellulose. Bioresource Technology, 72, 289294.
Larsson, S., Palmqvist, E., Hahn-Haagerdal, B., Tengborg, C.,
Stenberg, K., Zacchi, G., & Nilvebrant, N.-O. (1999). The
generation of fermentation inhibitors during dilute acid hydrolysis
of softwood. Enzyme and Microbial Technology, 24, 151159.
Lawford, H. G., & Rousseau, J. D. (1998). Improving fermentation
performance of recombinant Zymomonas in acetic acid-containing
media. Applied Biochemistry and Biotechnology, 7072, 161172.
Luo-Caidian, A., Brink-David, L., & Blanch-Harvey, W. (2002).
Identication of potential fermentation inhibitors in conversion of
hybrid poplar hydrolyzate to ethanol. Biomass and Bioenergy, 22,
125138.
Maiorella, B., Blanch, H. W., & Wilke, C. R. (1983). By-product
inhibition eects on ethanolic fermentation by Saccharomyces
cerevisiae. Biotechnology and Bioengineering, 125, 103121.
Maloney, M. T., Chapman, T. W., & Baker, A. J. (1985). Dilute acid
hydrolysis of paper birch: Kinetic study of xylan and acetyl-group
hydrolysis. Biotechnology and Bioengineering, 27, 355361.
Nigam, J. N. (1998). Single cell protein from pineapple cannery euent.
World Journal of Microbiology and Biotechnology, 14, 693696.
Palmqvist, E., Almeida, J. S., & Hahn-Haagerdal, B. (1999). Inuence of
furfural on anaerobic glycolytic kinetics of Saccharomyces cerevisiae
in batch culture. Biotechnology and Bioengineering, 62, 447454.
Paraj oo, J. C., Domnguez, H., & Domnguez, J. M. (1998). Biotech-
nological production of xylitol. Part 1: Interest of xylitol and
fundamentals of its biosynthesis. Bioresource Technology, 65, 191
201.
Paraj oo, J. C., Santos, V., & del Ro, F. (1995a). Hidr oolisis de la
fracci oon hemicelul oosica de la madera de pino. I. Cineetica y
distribuci oon de productos en operaci oon a presi oon atmosfeerica.
Anidad, 52, 162170.
Paraj oo, J. C., Santos, V., & del Ro, F. (1995b). Hidr oolisis de la
fracci oon hemicelul oosica de la madera de pino II. Operaci oon a
presiones superiores a la atmosfeerica. Anidad, 52, 267274.
Pessoa, A., Mancilha, I. M., & Sato, S. (1996). Cultivation of Candida
tropicalis in sugar cane hemicellulosic hydrolyzate for microbial
protein production. Journal of Biotechnology, 51, 8388.
Rivas, B., Domnguez, J. M., Domnguez, H., & Paraj oo, J. C. (2002).
Bioconversion of posthydrolysed autohydrolysis liquors: An alter-
native for xilitol production form corn cobs. Enzyme and Microbial
Technology, 31, 431438.
Rodrgues, C. G. A., Silva, S. S., Prata, R. P., & Felipe, A. (1998).
Biotechnological production of xylitol from agroindustrial resi-
dues. Applied Biochemistry and Biotechnology, 7072, 869875.
Saeman, J. F. (1945). Kinetics of wood saccharication. Hydrolysis of
cellulose and decomposition of sugars in dilute acid at high
temperature. Industrial and Engineering Chemistry, 37, 4352.
Teixeira, L. C., Linden, J. C., & Schroeder, H. A. (1999). Optimizing
peracetic acid pretreatment conditions for improved simultaneous
saccharication and co-fermentation (SSCF) of sugarcane bagasse
to ethanol fuel. Renewable Energy, 16, 10701073.
Teellez-Luis, S. J., Ramrez, J. A., & Vaazquez, M. (2002). Mathematical
modelling of hemicellulosic sugar production from sorghum straw.
Journal of Food Engineering, 52, 285291.
van Zyl, C., Prior, B. A., & Du Preez, J. C. (1991). Acetic acid
inhibition of D-xylose fermentation by Pichia stipitis. Enzyme and
Microbial Technology, 13, 8286.
Yahiro, K., Shibata, S., Jia, S.-R., Park, Y., & Okabe, M. (1997).
Ecient itaconic acid production from raw corn starch. Journal of
Fermentation and Bioengineering, 84, 375377.
152 A. Rodrguez-Chong et al. / Journal of Food Engineering 61 (2004) 143152

You might also like