Vous êtes sur la page 1sur 8

Bioresource Technology 46 (1993) 233-240

ON THE BEHAVIOUR OF LIGNIN A N D HEMICELLULOSES


DURING THE ACETOSOLV PROCESSING OF WOOD
J.C. Paraj6, J.L. A l o n s o & D. V~izquez

Department of Chemical Engineering, University of Vigo (Campus Orense), Las Lagunas, 32004 Orense, Spain
(Received 10 February 1993; revised version received 6 april 1993)

Abstract and lignin (untouched or as degradation products; see


Eucalyptus globulus wood samples were submitted to Davis et al., 1986; Young & Davis, 1986; Johansson et
fractionation with 95% acetic-acid solutions catalysed by al., 1987; Kin, 1990; Shukry et al., 1992; V~izquez et
small quantities of HCI. Experiments were performed at al., 1992), which can be convened by chemical or bio-
the normal boiling temperature using a liquor/wood technological means into a variety of chemicals.
ratio equal to lO g/g. The effects of treatment time (0-180 Acetic-acid media (with or without catalyst addition)
min) and catalyst concentration (0-0"2%) on pulp yield, have been employed for the fractionation of several
sofid residue composition and cooking liquor composi- lignocellulosics, including hardwoods, softwoods and
tion were determined. Pulp yields in the range bagasse (Davis et al., 1986; Nimz & Casten, 1986;
51.5-90"6% were obtained. Extensive delignification of Young & Davis, 1986; Kin, 1990; Shukry et al., 1992;
wood was achieved under a variety of operational condi- V~izquez et al., 1992). In the presence of small amounts
tions, leading to sofid residues with less than 4% lignin. of strong mineral acids (which act as catalysts), exten-
The lignin removal was modelled kinetically as the con- sive and selective delignification of raw materials can
tribution of two parallel, first-order reactions. Pulps con- be achieved by acetic acid solutions under mild experi-
taining over 90% polysaccharides and mental conditions. During the fractionation treatments,
glucan/polysaccharide ratios higher than 0"95 were the hemicellulose fraction is hydrolysed, whereas cellu-
obtained under selected conditions. Pentoses and lose remains as a solid residue.
furfural were identified as hemicellulose hydrolysis In an earlier work (V~quez et al., 1992), the authors
products. The kinetics of hemicellulose saccharification studied the delignification of Eucalyptus globulus by
was modelled assuming two sequential, first-order, HCl-catalysed acetic-acid solutions. The effects of
irreversible reactions. The maximum pentose recovery acetic-acid concentration, catalyst concentration and
accounted for 37% of the potential amount. Under the reaction time on the pulp yield and solid residue com-
conditions leading to optimum pentose recoveries, good position were assessed using empirical models. Under
delignification was achieved. Furfural concentrations up selected experimental conditions, eucalypt wood was
to 4"3 g/l were obtained under the severest experimental extensively delignified and the hemicelluloses were
conditions studied. selectively hydrolysed at pulp yields in the vicinity of
50%.
Keywords: Acetic acid pulping, organosolvent delignifi- Further studies demonstrated the possibility of
cation, acetosolv processing, hemicellulose hydrolysis. improving the above results (including both delignifica-
tion extent and pulp yield) using shorter reaction times
than those previously studied. Additionally, the influ-
INTRODUCTION ence of the operational conditions on the type and
amount of hemicellulose hydrolysis products was
The processing of lignocellulosic materials by included in this study, in order to provide an evaluation
organosolvents provides an interesting alternative to of the catalysed acetic-acid process as treatment for
conventional chemical processes. A variety of organo- biomass fractionation.
solvents (including alcohols, ketones, amines, glycols, This work deals on the optimization of the catalysed
esters, peroxides and acids) has been used to perform acetic-acid processing of Eucalyptus globulus wood.
the fractionation of biomass into its main constituents. The main objective was to provide further information
Organosolv-based processes allow the selective on the potentiality of acetic-acid-based processes for
separation and recovery of cellulose, hemicelluloses performing an extensive and selective separation of the
main biomass constituents, leading to fractions which
Bioresource Technology 0960-8524/93/S06.00 © 1993 could be used as chemical intermediates. Delignifica-
Elsevier Science Publishers Ltd, England. Printed in Great tion and hemicellulose hydrolysis were accomplished
Britain in a single step. Obtaining marketable products (xylose
233
234 J.C Paraj6, J.L. Alonso, D. V6zquez

or furfural) from hemicelluloses was considered. Fitting of data


Experiments were performed at the normal boiling The experimental data were fitted to theoretical
temperature using a liquor/wood ratio of 10 g/g. Since models by non-linear regression using commercial
high acetic acid concentration favours the overall pro- software (TableCurve from Jandel Scientific, Corta
cess (V~quez et al., 1992), 95% acetic acid solutions Madera, CA, USA).
were chosen as delignification media. The operational
variables considered were reaction time (in the range
0-180 min) and catalyst concentration (in the range RESULTS AND DISCUSSION
0-0"2% HC1). Pulp yield, solid residue composition
Raw material composition
and hemicellulose hydrolysis products (glucose,
pentoses and furfural) were considered as experimental The wood lot used for experimentation contained
variables. 23-3% Klason lignin and 62' 3% polysaccharides. The
glucan fraction of polysaccharides accounted for the
42.3 wt % of wood. These values are in agreement with
METHODS previously reported results (Paraj6 et al., 1992;
Vfizquez et al., 1992).
Raw material The hemicellulose fraction of eucalypt wood is
Eucalyptus globulus wood samples were collected in a mainly composed of xylose, followed by glucose and
local pulp mill (ENCE, Pontevedra), milled and arabinose. The HPLC columns used for analysis
screened to select the fraction of particles with a size separate glucose and arabinose as individual peaks,
between 0"25 and 1 mm. The wood particles were air- whereas galactose, mannose and xylose are eluted in a
dried, homogenised in a single lot to avoid differences single peak. Since the relative proportion of mannose
in composition, and stored. and galactose polymers in hemicelluloses is quite low
Analysis of wood (David et al., 1988) and the xylose/arabinose ratio is
Aliquots from the homogenised wood lot were sub- high (within the range 10-15 g/g in all the experiments
mitted to moisture determination and to quantitative performed), the authors have preferred to report all the
acid hydrolysis with 72% sulphuric acid following non-glucose, hemicellulosic sugars as pentoses; but it
standard methods (Browning, 1967; Vfizquez et al., must be taken into account that at least 90% of these
1987). The solid residue after hydrolysis was con- pentoses were xylose.
sidered as Klason lignin. The monosaccharides con-
tained in hydrolysates were determined by Operational conditions used in fractionation
spectrophotometric or HPLC methods (see below). experiments
In an earlier work (Vfizquez et al., 1992), the authors
Processing of wood samples studied the delignification of Eucalyptus globulus wood
Wood chips and acetic-acid-water solutions were by HCl-catalysed acetic acid solutions. Empirical
placed in Erlenmeyer flasks and heated to the normal models were developed to assess the effects of acetic
boiling temperature. Then, concentrated HC1 solutions acid concentration (within the range 75-95%), HCI
were added to the desired concentration and the flasks concentration (0-0"2%) and reaction time (5-6 h) on
were maintained under reflux at atmospheric pressure. both pulp yield and solid residue composition. All the
The liquor/wood ratio used in experiments was 10 g/g. experiments were carried out at the normal boiling
It was assumed that both prehydrolysis and delignifica- point of solutions using liquor/wood ratio of 10 g/g.
tion started when the catalyst was added. The moisture T h e experimental variables depended mainly on the
content of wood (10%) was considered as water in the catalyst concentration used, whereas high acetic acid
material balances. At the end of delignification, the concentrations improved the reaction selectivity. The
solid residue was recovered by filtration, washed with experimental variables were not significantly affected
acetone, air-dried and submitted to quantitative by the reaction time in the range considered. Using
saccharification. The pulping liquors were mixed with 0.15% HC1 and 95% acetic-acid solutions, the models
water (1:10) to precipitate the lignin fraction, and the predicted solid residues with 6"6% lignin, 87.4% poly-
solid residues were removed by centrifugation. saccharide and 0-928 glucan/polysaccharide ratio at
Aliquots from supernatants were filtered through 0.45 45% pulp yield.
/~m membranes and analysed by HPLC. Additional studies demonstrated that several factors
involved in the fractionation process could be
HPLC analysis of monosaccharides and furfural improved by using shorter reaction times, including
Samples from quantitative acid hydrolyses or from pulp yields, degree of delignification and recovery of
pulping liquors were analysed for glucose, pentoses valuable hemicellulose hydrolysis products.
(xylose and arabinose) and furfural using a Waters This paper reports on the additional work per-
HPLC with refractive-index detector. The operational formed to provide further insight on the chemical pro-
conditions were: columns Shodex SH1011 (two); cessing of eucalypt wood using HCl-catalysed acetic
mobile phase, H2SO 4 0"015 m; column temperature, acid solutions. Experiments were performed using the
90°C; flow rate, 1 ml/min. same temperature and liquor/wood ratio previously
A cetosolv processing of wood 235

studied (V~quez et aL, 1992). In order to improve the reaction time accounted for over 86% of the values
selectivity of fractionation, delignification media con- determined for samples treated for 180 min.
taining 95% acetic acid were used in all the cases con- For a given reaction time, an increase in catalyst con-
sidered. The operational variables studied were centration caused a decrease in pulp yield. This behav-
catalyst concentration (in the range 0-0.2% HCI) and iour is in agreement with the data obatined for
reaction time (0-180 min). Pulp yield, solid residue delignification and carbohydrate hydrolysis (see
composition and cooking liquor composition were below). Significant decreases in pulp yield were
measured. Figure 1 summarises the whole processing observed when the catalyst concentration was
of samples, including both fractionation and analysis increased. In experiments lasting 180 min, an increase
stages. in HCI concentration from 0"05 to 0-10% caused a
marked decrease in yield (from 62.3 to 55.4%, see
experiments 6 and 12 of Table 1 ); but when the catalyst
Pulp yield concentration was increased from 0"15 to 0-20%,
The weight loss of wood during fractionation is minor variations in pulp yield were observed (from
mainly caused by the combined effect of extractives 52.4 to 51"5%, see experiments 18 and 24 of Table 1).
removal, delignification and carbohydrate hydrolysis. This fact suggests that under the experimental condi-
Table 1 shows the experimental conditions assayed and tions assayed, a practical limit for yield in the vicinity of
the results obtained. Pulp yields in the range 50% exists.
90-6-51" 5% were obtained in the various experiments. In comparison with reported results, lower yields
Even in the experiments performed under the severest (48-48" 5%) were obtained for eucalypt wood in treat-
operational conditions studied (120-180 min reaction ments with HCl-catalysed acetic acid lasting 5-6 h
time, 0" 15-0.20% catalyst concentration), good pulp (V~izquez et aL, 1992). The acetic-acid pulping of
yields (within the range 51 "5-53.9%) were obtained. woods from other tree species (such as beech, aspen or
The kinetics of weight loss followed general trends spruce) gave yields in good agreement with the values
similar to reported results (Young & Davis, 1986; found in this work (Davis et aL, 1986; Young & Davis,
Davis et aL, 1986). The variation pattern of data listed 1986; Kin, 1990).
in Table 1 suggests two sequential stages with different
rates, the first being faster than the second. In experi- Lignincontentanddelignificationkinetics
ments performed with catalyst concentrations higher Table 1 lists the lignin content of solid residues
than 0" 10%, the weight loss observed after 60 min of obtained under the conditions studied. It can be noted

Wood s a m p l e s

Acetic acid/ ~D
H20/HCI ELIGNIFICATIO
~ Suspension
Liquors

Solid residue ~ FILTRATIONI


Acetone ~ WASHING
I ~ Washing liquors ILIGNIN PRECIPITATION ]4 Water
r Solid r e s i d u e Suspension

~i~quotAIR DRYING~----~ tl2so~ I CENTR|FUGATION I


1 Aliquot 2~

(To m o i s t u r e
and yield
]SACCHARI
[
FICATI/~-~]
ON Liquors
Solid residue Supernatant
determination ) (lignin) (to m o n o s a e c h a r i d e
Solid I ~' and furfural d e t e r m i n a t i o n )
residue~ ~
NEUTRALIZATION Ba(OH) 2

J liquot 1
[OVENDRYING] ~r ,
(To m o n o s n c c h a r i d e d e t e r m i n a t i o n )

r~_.~ Aliquot 2 (To glucose determination)


K l a s o n Iignin
Fig. 1. Scheme of the experimental work performed.
236 J.C. Paraj6, J.L. Alonso, D. V6zquez

Table 1. Operational conditions studied and experimental results obtained for yield and pulp composition

Experiment Operational variables Experimental variables

HCI (%) Time (min) Yield (%) Lignin (%) Polysaccharide (%) GPR"

1 0.05 15 90"6 18"3 67.2 0.758


2 0.05 30 81.7 16.1 74'1 0.794
3 0-05 60 73"2 15"5 74"8 0.820
4 0.05 90 69-1 13"0 80-0 0.874
5 0.05 120 62.7 8"3 83"2 0.888
6 0.05 180 62.3 6"8 82-1 0.881
7 0-10 15 86'5 17"7 69"2 0"800
8 0.10 30 66"8 13-3 79'3 0"801
9 0.10 60 61.1 9"6 84"8 0.859
10 0.10 90 56.8 8"2 88.0 0-892
11 0.10 120 55.8 6.1 91.1 0"910
12 0-10 180 55-4 5.0 89"0 0"901
13 0.15 15 73.5 13-7 73"6 0.836
14 0.15 30 63"3 10.7 85"6 0"896
15 0.15 60 57.8 6"2 86.6 0.906
16 0-15 90 55"1 5"6 89.7 0.948
17 0.15 120 53-9 3.7 91.3 0-951
18 0.15 180 52.4 3'9 90"7 0"926
19 0.20 15 72"5 13"6 73.8 0"842
20 0.20 30 62.2 8"8 84.9 0"888
21 0.20 60 56.2 4.4 90.2 0-917
22 0-20 90 54.9 3-1 89.1 0"948
23 0.20 120 53.4 3"3 90-8 0"964
24 0.20 180 51"5 3'5 89.4 0"911
"GPR, Glucan/polysaccharide ratio.

that extensive delignification can be achieved under a the data of Table 1 can be correlated according to the
variety of experimental conditions. The lowest lignin following scheme:
content obtained (3" 1 wt % of solid, experiment 22) is Lf
less than 50% of the minimum lignin content reported
in our previous work using longer reaction times Ld
(V~izquez et al., 1992). This fact is thought to be caused Z
Ls
by lignin repolymerization reactions. In fact, the
analysis of data corresponding to experiments 17, 18, Assuming first-order kinetics for both reactions, it can
22, 23 and 24 suggests that repolymerisation occurred be shown that:
even under the mild conditions considered in this
work. Extensive repolymerisation has been reported in P R L = y L C =a'e-k~'+ (1 - a)e -k~'
acid-catalysed delignification of beechwood (Kin, LCo
1990).
The kinetic pattern of delignification is closely where PRL is the percentage of residual lignin in wood,
related to that of weight loss previously discussed. An Y is the pulp yield; LC and LCo are the lignin contents
initial, fast stage (bulk delignification) is followed by a of treated and untreated wood, respectively; a is the
slower residual delignification stage. Similar behaviour fraction of fast-reacting lignin (defined as L f/[L f + L s]);
has been reported for both Kraft- and organosolv- t is the reaction time, and kf and k s are first-order
processing of wood (Cho & Sarkanen, 1985; Tirtowi- kinetic coefficients for the fast- and slow-reacting lignin
djojo et al., 1988). An interpretation of the fractions.
experimental results can be provided by a kinetic Table 2 lists the values of kf, k s and a obtained by
model assuming the presence in wood of two lignin nonlinear regression of a PRL/time series of data
fractions, a fast-reacting fraction (L f) and slow-reacting corresponding to a given HCI concentration, as well as
fraction (Ls), which are depolymerised at different the values obtained for the statistical parameters ra and
rates to give soluble lignin-fragments (L0); whereas the F measuring the correlation and significance of models.
depolymerised lignin L 0 can undergo repolymerisation Figure 2 allows a comparison between experimental
reactions to give insoluble, repolymerised lignin (Lr). results of PRL (calculated from the data listed in Table
Since the repolymerisation effects have a limited 1) and the model predictions for the various catalyst
importance under the experimental conditions studied, concentration used. The close agreement between
Acetosolv processing of wood 237

/
Table 2. Regression and statistical parameters obtained in
the modelling of wood delignification qgO
tu
>.
HCl a kf ks r2 F
concentration (min- ]) (min- 1) ~, 80.
(%) rl

70
0"05 0'244 0"120 0"008 0"988 166
...'"
0"10 0"765 0"043 0"004 0"995 410
0"15 0"756 0"080 0"007 0"997 746 60
ii~ ' A " . . J l , , r•.
0"20 0"908 0"065 0"001 0"999 2527
50
0 2 4 6 8 10 12 14 16 18 20

L I G N I N C O N T E N T (%)

Fig. 3. Interrelationship between pulp yield and operational


variables (A, 0.05% HC1; o, 0'10% HCI; n, 0-15% HCI; Y,
_z 80 0.20% HCI)
Z
L~
_J
~ 60
-'1
a importance on the polysaccharide content. Samples
uJ with polysaccharide contents in the vicinity of 91%
r~ 40
were obtained in experiments performed with
ZI..- 20 O 0-10-0"20% HCI and 120 min reaction time. When
uJ ,
0 : the reaction time was increased from 120 to 180 min,
w i
small decreases in the polysaccharide content were
0 20 40 60 80 100 120 140 160 180 observed. The lower polysaccharide content reported
TIME (min) for delignification treatments lasting 5-6 h (87-5-84" 5)
Fig. 2. Experimental and predicted dependence of the can be explained in part by lignin repolymerisation
percent of residual lignin in wood samples (PRL) on the reactions.
operational variables studied (A ,0"05% HCI; o, 0.10% HCI; The glucan/polysaccharide ratio (GPR) of solid
s, 0.15% HC1; , , 0.20% HCI) residues provides information about the behaviour of
hemicelluloses during the chemical processing of
wood. The value of GPR determined for untreated
wood (0-675) increased to values in the range
experimental and predicted data confirms the reliabil- 0"758-0"964 for delignified solid residues. Both the
ity of the above equation in describing the whole kinetics and the extent of hemicellulose hydrolysis
delignification process. increased when higher catalyst concentrations were
Figure 3 shows the interrelationship between pulp used (see Table 1). The GPR values increased steadily
yields and lignin contents of samples ~treated under the with time in the range 0-120 min for all the catalyst
conditions listed in Table 1. Experiments performed concentrations assayed. Reaction times higher than
using 0-05% catalyst show a different behaviour to 120 min resulted in decreased values of GPR. This
those carried out with 0" 10-0"20% catalyst. Pulps with behaviour, which is difficult to explain from a theoreti-
6"8% residual lignin can be obtained at 62-3% pulp cal viewpoint, could be related to reversion reactions of
yield using low HCI concentrations (see experiment 6), sugars and/or experimental errors. In fact, glucan
whereas a similar lignin content was obtained at shorter recoveries after fractionation slightly higher than 100%
reaction times in trials performed with 0-10% catalyst, can be calculated from the experimental results.
but with decreased pulp yield (55.8% in experiment Reported results on the acetic treatment of wood are in
11). good agreement with the values of GPR determined in
this work (Kin, 1990; V~izquez et al., 1992).
Polysaccharide and glucan content
The polysaccharide content of solid residues from Hemiceilulose hydrolysis
the fractionation process depended on two factors: The recovery of hemicellulose reaction products as
delignification (which caused an enrichment of treated valuable chemicals (such as sugars or furfural) is an
samples in polysaccharides) and carbohydrate- important factor affecting the economic feasibility of
hydrolysis reactions (mainly hemicellulose hydrolysis) biomass fractionation processes. The HPLC analysis of
which led to pulps with improved cellulose contents. pulping liquors (after precipitation of the lignin fraction
The data listed in Table 1 show that for a given by water addition) provided useful information on the
reaction time significant changes in polysaccharide type and amount of marketable compounds generated
content were obtained when the catalyst concentration from hemicelluloses.
was increased from 0-05 to 0" 1%. Further increases in Xylose was the main product of hemicellulose
catalyst concentration caused effects of decreasing hydrolysis. Other sugars (arabinose and glucose) were
238 J.C. Paraj6, J.L. Alonso, D. V~zquez

also identified in chromatograms, but their relative model (Saeman, 1945):


concentrations were low. Since the resolution of xylose Pentosan ~ Pentoses ~ Decomposition products
and arabinose was poor owing to the large proportion
of xylose, both compounds were integrated as a single The equation giving the dependence of the pentose
peak, and the whole amount determined is referred to concentration (Cp) with time derived from the above
as 'pentoses'. The glucose concentrations determined reactions assuming first-order kinetics, is:
in the various experiments performed (0-0.63 g/l)
were not significant for the objectives of this study.
In experiments performed under strong operational
Cp=CpP'k~kl " ( e - k l / - - e

conditions, significant amounts of furfural (produced


by pentose decomposition) were observed. Furfural has where Cpp is the potential pentose concentration corre-
been previously reported to be a byproduct of the sponding to theoretical conversion of pentosan into
acetolysis of beechwood (Kin, 1990). The pattern of sugars, calculated from material balances; and k~ and
hemicellulose degradation (pentosan hydrolysis to k 2 are the first-order coefficients for pentosan hydroly-
sugars, followed by sugar decomposition) is the same sis and pentose degradation, respectively.
as that reported for the acid-catalysed hydrolysis of Table 4 lists the kinetic coefficients obtained by
hemiceiluloses in aqueous media (Conner, 1984; regression as well as the r 2 and F statistical parameters.
Ranganathan et al., 1985; Maloney et al., 1986; Naka- Figure 4 shows the agreement between experimental
gawa et al., 1986; Kim & Lee, 1987), but the reaction and calculated data of pentose concentration. The
rate of the reaction was quite different in of the present maximum pentose concentration (Cpmax)can be calcu-
authors' case. lated using the equation:
Table 3 includes the experimental values of pentose
concentration determined under the conditions kll(k2/k2-kl)
assayed. In all the series of data, the pentose concentra- Cpmax=Cpp'~k22)
tion reach a maximum. Both the maximum pentose
concentration and the time necessary to achieve this The values of Cpmaxcalculated from the data listed in
maximum depended on the catalyst concentration Table 4 varied in a narrow range (7"3 - 7"5 g/l), corre-
used. Since this behaviour was in qualitative agreement sponding to 3 6 . 5 - 37.5% of the potential concentra-
with the kinetic pattern described above, the data were tion (?pp.
fitted according to the well known Saeman's kinetic

Table 4. Kinetic coefficients and statistical parameters


obtained in the modelling of hemiceilulose hydrolysis
Table 3. Pentose concentration (Cp) determined in pulping
liquors HCI k1 k2 R2 F
concentration (min- t) (min- l)
Experiment Cp
(g/litre) 0-05 0.0050 0-0051 0-976 200
0-10 0.0130 0"0132 0"945 86
1 1"74 0"15 0.0201 0-0202 0"929 65
2 1"81 0"20 0.220 0-0210 0.976 206
3 4"32
4 5"88
5 7"11
6 6-94
A
7 1"83
8 6"10 d~
9 7-11 Z 6-
10 7"22 _o
11 6-62
12 4"23 z
LtJ 4-
0
13 5"79 Z
C,
14 6"02
15 6"75
16 5"78
17 4-74
18 2"20
19 4"79 0 ~

0
i

20
i

40
i

60 80
i ,

100
i , i

120
, i

140
• i

160 180
20 7"32 TIME (rnin)
21 6"72
22 5-33 Fig. 4. Experimental and predicted dependence of the
23 4"38 pentose concentration in pulping fiquors (C_) on temperature
"
24 2"10 and catalyst concentrataon ( • ,0"05 0YoHCI,• Pe, 0"10 0YoHCI,• m,
0.15% HCI; T, 0"20% HCI)
A cetosoIv processing of wood 239

Under the experimental conditions which led to It can be observed that the whole amount obtained by
maximum pentose concentration, samples with addition of the pentose concentration in liquid phase to
80-88% delignification can be obtained (see Fig. 2). the potential pentose concentration corresponding to
This compatibility of conditions for hemicellulose the pentosan remaining in solid phase and the concen-
hydrolysis and delignification reactions is a favourable tration of pentoses necessary to give the furfural con-
feature of the studied process. centration observed assuming theoretical yield, is far
Furfural was the main product of pentose degrada- below the value of the potential pentose concentration
tion. Figure 5 shows the dependence of the furfural Cpp in all the experiments performed (see Table 5).
concentration on the operational variables considered. This finding can be caused by several factors such as
formation of water-soluble xylose oligomers, genera-
tion of reversion products from sugars in solution (in
5 the same way as found for carbohydrate hydrolysis in
--I
aqueous media, see Mok et al., 1992), formation of
z
0
4 condensation products from furfural, and generation of
acetic-acid soluble, water-insoluble xylose polymers
rr
I--
z
3 which are associated with the phenolic fraction and
uJ
:z
which coprecipitate with lignin after water addition
0 2 (see Fig. 1 ).
el-

n-"
ACKNOWLEDGEMENT
o
20 40 60 80 100 120 140 160 180 The authors are grateful to 'Xunta de Galicia' for the
TIME (rain) financial support of this work (Proj. X U G A 3 8 3 0 2
Fig. 5. Influence of temperature and catalyst concentration A91).
on the furfural concentration (& ,0-05% HCI; o, 0" 10% HCI;
", 0" 15% HCI; I1', 0"20% HC1)
REFERENCES

Table 5. Pentosan and pentosan reaction products deter- Browning, B.L. (1967). Methods of Wood Chemistry. John
mined in experiments (results expressed as percent of the Wiley, New York.
potential pentose concentration Cpp) Cho, HJ. & Sarkanen, K.V. (1985). Alternatives to H-factor
measurements in the Kraft process. Paperi ja Puu, 67,
Experiment Treatment liquors Solid Total 121-4.
residue Conner, A.H. (1984). Kinetic modeling of hardwood pre-
Pentoses Furfural a Pentosan ~ hydrolysis. Part I. Xylan removal by water prehydrolysis.
Wood Fiber Sci., 16, 268-277.
1 8.7 0 73"6 82"3 David, C., Fornassier, R., Lejong, W. & Vanlautem, N.
2 9.1 0 62.4 71.5 (1988). Pretreatment of Eucalyptus wood with sodium
3 21.6 0 46"8 68.4 hypochloryte and enzymatic hydrolysis with cellulases of
4 29.4 0 34"9 64.3 Trichoderma viride. J. Appl. Polym. Sci., 36, 29-41.
5 35-6 0 29"2 64.8 Davis, J.L., Young, R.A. & Deodhar, S.S. (1986). Organic
pulping of wood. IlL Acetic acid pulping of spruce.
6 34.7 7.3 30.4 72.4 Mokuzai Gakkaishi, 32, 905-14.
7 9.2 0 59"9 69.1 Johansson, A., Aaltonen, O. & Ylinen, P. (1987). Organosolv
8 30"5 4.8 52"7 88"0 lignin pulping. Methods and pulp properties. Biomass, 13,
9 35"6 9.1 36"5 81.2
10 36"1 12.6 27"0 75.7 45-65.
Kim, S.B. & Lee, Y.Y. (1987). Kinetics in acid-catalyzed
11 33.1 13.1 22.8 69"0 hydrolysis of hardwood hemicellulose. Biotechnol. Bioeng.
12 21.2 17.6 24.4 63-2 Symp., 17, 71-83.
13 29.0 8.0 44.4 81.4 Kin, Z. (1990). The acetolysis of beech wood. Tappi J., 11,
14 30.1 6"8 28.2 65.1 237-238.
15 33"8 14.3 23.6 71.7 Maloney, M.T., Chapman, T.W. & Baker, AJ. (1986). An
engineering analysis of the production of xylose by dilute
16 28.9 19-7 12.8 61.4 acid hydrolysis of hardwood hemicellulose. Biotechnol.
17 23.7 24.5 12.1 60.3 Prog., 2, 192-202.
18 11"0 32.0 17.6 60-6 Mok, W.S.L., Antal, MJ. & Varhegyi, G. (1992). Productive
19 24.0 5"5 42.3 71.8 and parasitic pathways in dilute acid-catalyzed hydrolysis
20 36.6 8"7 29.6 74.9
of cellulose. Ind. Eng. Chem. Res., 31, 94-100
21 33.6 20"6 21.1 75"3 Nakagawa, M., Kamiyama, Y. & Sakai, Y. (1986). Saccharifi-
22 26.7 23.7 12.7 63"1 cation of tropical wood. I. Hydrolysis rates of giant
23 21.9 28.2 8.8 58-9 ipil-ipil wood pentosan and cellulose in dilute sulfuric
24 10.5 33"5 20-5 64.5 acid. Japanese J. Tropical Agric., 30, 153-9.
Nimz, H.H. & Casten, R. (1986). Chemical processing of
aFurfural and pentosan expressed as pentose equivalents. lignocellulosics. Holz Roh und Werk., 44, 207-212.
240 J.C. Para]6, ZL. Alonso, O. V6zquez

Paraj6. J.C., Alonso, J.L., Lage, M.A. & V~quez, D. (1992). Tirtowidjojo, S., Sarkanen, K.V., Pla, F. & McCarthy, J.L.
Empirical modeling of Eucalyptus wood processing. Bio- (1988). Kinetics of organosolv delignification in batch and
proc. Eng., 8, 129-136. flow-through reactors. Holzforschung, 42, 177-183.
Ranganathan, S., McDonald, D.G. & Bakhshi, N.N. (1985). V~quez, G., Antorrena, G. & Paraj6, J.C. (1987). Studies on
Kinetic studies of wheat hydrolysis using sulphuric acid. the utilization of Pinus pinasterbark. 1. Chemical constitu-
Can. J. Chem. Eng., 63, 840-844. ents. Wood Sci. Technol., 2 l, 65-74.
Saeman, J.F. (1945). Kinetics of wood saccharification. Ind. V~quez, D., Lage, M.A., Paraj6, J.C. & V~quez, G. (1992).
Eng. Chem., 37, 43-52. Fractionation of Eucalyptus wood in acetic acid media.
Shukry, N., E1-Meadawy, S.A. & Nassar, M•. (1992). Pulp- Biores. Technol., 40, 131-136.
ing with organic acids: 3-acetic acid pulping of bagasse. J. Young, R.A. & Davis, J.L. (1986). Organic acid pulping of
Chem. Tech. Biotechnol., 54, 135-143. wood. Part II. Acetic acid pulping of aspen. Holzforschung,
40, 99-108.

Vous aimerez peut-être aussi