Вы находитесь на странице: 1из 14

Chemical Engineering Journal 171 (2011) 883–896

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Dye adsorption on autohydrolyzed pine sawdust in batch and fixed-bed systems


D. Sidiras a,∗ , F. Batzias a , E. Schroeder b , R. Ranjan c , M. Tsapatsis c
a
Laboratory of Simulation of Industrial Processes, Department of Industrial Management and Technology, University of Piraeus, 80 Karaoli & Dimitriou, GR 18534 Piraeus, Greece
b
Institut fuer Kern- und Energietechnik, Forschungszentrum Karlsruhe GmbH, Postfach 3640, 76021 Karlsruhe, Germany
c
Department of Chemical Engineering and Materials Science, University of Minnesota, 421 Washington Avenue SE, Minneapolis, MN 55455, USA

a r t i c l e i n f o a b s t r a c t

Article history: Batch and column kinetics of three basic dyes, methylene blue (MB), bismarck brown (BB) and acri-
Received 13 January 2011 dine orange (AO), adsorption on autohydrolyzed (160–240 ◦ C, 0–50 min) Scots pine (Pinus sylvestris L.)
Received in revised form 13 April 2011 sawdust were studied, using untreated pine sawdust as control, in order to explore the potential use
Accepted 14 April 2011
of these low-cost industrial byproducts (untreated autohydrolyzed sawdust) for wastewater cleaning.
The effect of the autohydrolysis time and temperature on the microstructure and the crystallinity of
Keywords:
pine sawdust was investigated be means of: SEM, FTIR, XRD and BET, in order to obtain information
Adsorption
at a higher granularity level, in comparison with previous studies. The BET surface area increased from
Autohydrolysis
Dye
0.89 to 19.3 m2 g−1 . In the case of MB, the Freundlich’s adsorption capacity KF increased from 5.60 to
Sawdust 15.7 (mg g−1 )(L mg−1 )1/n , the amount of dye adsorbed when saturation is attained (Langmuir constant qm )
Column increased from 38.7 to 88.0 mg g−1 , and the Bohart-Adams adsorption capacity coefficient N increased
from 8046 to 14157 mg L−1 , indicating the extent that the severity of the autohydrolysis treatment
enhanced the adsorption behavior of the material.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction for the removal of dyes in aqueous solutions. The environmental


impact of the dyes present in the wastewater streams of many
Autohydrolysis of lignocellulosic materials like sawdust is industrial sectors, such as textile, paper, dyeing, tannery and the
already used in bioethanol industry for water soluble fermentable paint industry, has drawn a lot of attention, emphasizing the neces-
sugars production [1,2]. The main idea in this manuscript is the sity for their removal. In particular, wood sawdust [9], a relatively
use of the autohydrolysis solid residue as an adsorbent. In fact, the abundant and inexpensive material, has been extensively inves-
adsorbent is the solid by-product in the fermentable sugars produc- tigated as an adsorbent for removing contaminants from water.
tion by autohydrolysis. Moreover, sawdust is a by-product of the Other adsorbent materials that have been studied include indus-
wood industry and pure water is used as autohydrolysis reagent trial byproducts and agricultural residues, untreated or pretreated
can be recycled after the fermentation and the water/bioethanol in various ways.
distillation processes. Consequently, the present work might also Among the untreated materials that have been investigated,
be considered within an Industrial Ecology framework. many agricultural residues, such as wheat straw, rice husk, corn-
Autohydrolysis of lignocellulosic materials can be used as a cobs and wood chips, have been used successfully to adsorb
first step while enzymatic or acid hydrolysis is usually used as individual dyes and dye mixtures in textile effluent [10]. Removal of
second step in the bioethanol industry. Autohydrolysis might sig- MB and other basic dyes has been carried out using beech sawdust
nificantly increase the enzymatic hydrolysis efficiency [1]. Many [9,11], wheat straw [12], cedar sawdust [13], rubberwood sawdust
industrial byproducts (e.g. sawdust, and wood-chips) and agricul- [14], kudzu [15], banana and orange peels [16] and palm kernel
tural residues, e.g. corncobs, almond shells, olive stones, rice husks, fiber [17].
wheat straw, and barley straw, can be used as feedstock for the Acid and alkali pretreated lignocellulosic materials (wheat
production of xylo-oligosaccharides by autohydrolysis [2]. straw, corncobs, barley husks, wood sawdust) were successfully
According to many review papers [3–8] low-cost adsorbents used as adsorbents for a variety of dyes [18–20]. Prehydrolysed
offer a lot of promising benefits for commercial purposes in the (with dilute sulphuric acid aquatic solution at 100 ◦ C) wheat straw
future. They could be used in place of commercial activated carbon [12] and beech sawdust [21] and chloride salts treated (at 100 ◦ C)
beech sawdust [9,11] has been proven to be effective for basic dyes
adsorption in batch and fixed-bed systems.
∗ Corresponding author. Tel.: +30 2104142360; fax: +30 2104142392. The autohydrolysis was chosen as a pretreatment (a) because
E-mail address: sidiras@unipi.gr (D. Sidiras).
compared with the other treatments with aquatic solutions uses

1385-8947/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2011.04.029
884 D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896

Nomenclature
T␪a autohydrolysis temperature in ◦ C
AO acridine orange u linear velocity (cm min−1 ) in the case of column
BB bismarck brown adsorption process
BET Brunauer, Emmet and Teller x bed depth of the adsorption column (cm)
bKF empirical parameter of eq. (4) XRD X-ray diffraction
bqm empirical parameter of eq. (8)
C concentrations of MB or BB or AO in the bulk solution  XRD angle (◦ )
at time t in the case of batch adsorption process;  order of the adsorption kinetic model
also, the effluent concentration (mg L−1 ) in the case  spectrophotometer wavelength (nm)
of the column adsorption process
c intercept of the intra-particle diffusion equation
(mg g−1 )
C0 initial dye concentration in the case of batch adsorp- only pure water as a reagent, and (b) this treatment is already used
tion process (mg L−1 ) as first step in the fermentable sugars for bioethanol production
Ce equilibrium concentration of the adsorbate (mg L−1 ) industry [1,2].
for t→ ∞ The suitability of a range of materials as dye adsorbents is usu-
Ci influent concentration (mg L−1 ) in the case of col- ally established using the kinetics of MB (or other dye) adsorption
umn adsorption process during batch and continuous (column) processes [22–26].
E adsorption activation energy (kJ mol−1 ) In this work, the removal of MB, BB and AO by autohydrolyzed
FTIR Fourier transform infrared pine sawdust was studied using untreated pine sawdust as con-
K adsorption rate coefficient (L mg−1 min−1 ) in the trol. The effect of the autohydrolysis temperature and time on the
case of column adsorption process microstructure and the crystallinity of pine sawdust was deter-
k first order rate constant for the batch adsorption mined be means of: SEM, FTIR, XRD and BET, in order to obtain
process (min−1 ) information at a higher granularity level, in comparison with pre-
k2 second order rate constant for the batch adsorption vious studies. The batch and column adsorption kinetics of these
process (min−1 ) dyes were used to estimate and compare the adsorption capac-
KF Freundlich constant related to adsorption capacity ity of the untreated and treated pine sawdust. The contribution of
(mg g−1 )(L mg−1 )1/n this paper concerns the thorough investigation of the effect of pure
KF,0 empirical parameter of eq. (4) water treatment at 160–240 ◦ C on sawdust. The autohydrolysis
KL Langmuir constant related to the energy of adsorp- treated sawdust in this temperature range simulates the byproduct
tion (L mg−1 ) in the fermentable sugar production, part of the second generation
kp intra-particle diffusion rate constant bioethanol producing industry.
(mg g−1 min−0.5 )
k␬ -order rate constant for the batch adsorption pro- 2. Materials and methods
cess (min−1 ).
m weight of the adsorbent used (g) 2.1. Adsorbent processing
MB methylene blue
N adsorption capacity coefficient (mg L−1 ) in the case The Scots pine (Pinus sylvestris L.) sawdust used was obtained
of column adsorption process from a local furniture manufacturing company, as a suitable source
n inverse of the slope of the Freundlich isotherm, it is for full-scale/industrial applications. The moisture content of the
related to adsorption intensity material when received was 8.7% (w/w); after screening, the frac-
NLRA non-linear regression analysis tion with particle sizes between 0.2 and 1 mm was isolated. The
p frequency factor at the Arrhenius law (min−1 ) composition of the raw material was as follows (expressed in % w/w
q amount adsorbed per unit mass of the adsorbent for on a dry weight basis): 40.1% cellulose measured as glucan (with
t→ ∞ (mg g−1 ) 52.5% XRD degree of crystallinity); 28.5% hemicelluloses (16.0%
qm Langmuir constant related the amount of dye measured as manan, 8.9% measured as xylan and the rest 3.6% mea-
adsorbed (mg g−1 ) when the saturation is attained sured as arabinan); 27.7% Klason acid-insoluble lignin, 0.2% ash,
qm,0 empirical parameter of eq. (8) and 3.5% extractives and other acid soluble components (e.g. acid
qt amounts of dye adsorbed per unit mass of the adsor- soluble lignin).
bent (in mg g−1 ) at time t The autohydrolysis process was performed in a 3.75-L batch
R2 coefficient of determination reactor PARR 4843. The isothermal autohydrolysis time was tai = 0,
RL dimensionless constant called ‘equilibrium param- 10, 20, 30, 40 and 50 min (not including the non-isothermal pre-
eter’ or ‘separation factor’ expressing the essential heating and the cooling time-periods); the reaction was catalyzed
characteristics of the Langmuir isotherm by the organic acids produced by the pine sawdust itself during
SEE standard error of estimate autohydrolysis at a liquid-to-solid ratio of 10:1; the liquid phase
SEM scanning electron microscope volume (water) was 2000 mL and the solid material dose (pine
t adsorption time (in min) sawdust) was 200 g; stirring speed 150 rpm. The reaction ending
T adsorption temperature in K temperature values were T␪a = 160 ◦ C, 200 ◦ C and 240 ◦ C, reached
ta autohydrolysis time (in min), ta = tp + ta after tp = 42, 62 and 80 min preheating time values, respectively.
Ta autohydrolysis temperature in K. The autohydrolysis product was filtered using a Buchner filter
tai autohydrolysis isothermal time (in min) with Munktell paper sheet (grade 34/N) to separate the liquid
tp autohydrolysis preheating time (in min) phase and from the solid phase. The solid residue was washed
T␪ adsorption temperature in ◦ C with water until neutral pH (the initial filtrate pH was 2.90–4.76
depending on the autohydrolysis severity). The solid residue was
D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896 885

1.5–13 (the initial pH of the dye solutions was adjusted using dilute
H2 SO4 or NaOH solutions, as appropriate).

2.4. Column studies

Fixed-bed up-flow adsorber studies were conducted in


15 cm × 2.5 cm and 25 cm × 2.5 cm stainless steel columns; the bed
height was x = 15 cm and 25 cm, respectively. The experimental
set-up consisted of one parallel column, fed by a multi-channel
peristaltic pump at constant flow rates Q = 10, 20, 30, 40 mL min−1 .
The initial constant MB concentration was at the ranges of
Ci = 14.0–14.3 mg L−1 and 156–192 mg L−1 . The initial constant BB
concentration was at the range of Ci = 11.0–11.1 mg L−1 . The initial
constant AO concentration was at the range of Ci = 8.9–9.0 mg L−1 .
Interconnective tubing and fittings were made of polytetrafluo-
roethylene (PTFE). Effluent samples were analyzed to yield output
concentration breakthrough curves.

2.5. Analytical techniques and instruments

The degree of crystallinity of pine sawdust cellulose was


measured by XRD [27] and by acid hydrolysis [21]. The resisting-
Fig. 1. (a) Solid residue yield of the pretreated sawdust and (b) lignin percentage of to-reaction hemicelluloses fraction (mainly xylan 1) was also
this residue vs. autohydrolysis time. measured by acid hydrolysis [21]. The Saeman et al. [28] technique
was used for the quantitative saccharification of the untreated
lignocellulosic material and the autohydrolysis reaction solid
dried at 110 ◦ C for 10 days at room temperature to reach the residues. The filtrates from the quantitative saccharification, were
humidity of the untreated material. Then it was used as adsor- analyzed for glucose, xylose and arabinose using high-performance
bent. liquid chromatography (HPLC, Agilent 1200) with Aminex HPX-
87H Column, refractive index detector and 5 mM H2 SO4 in water
as the mobile phase. Cellulose was estimated as glucan and hemi-
2.2. Adsorption isotherm studies celluloses were estimated as xylan and arabinan. Finally, the
acid-insoluble lignin (Klason lignin) was determined according to
Adsorption isotherms were derived from batch experiments. the Tappi T222 om-88 method [29].
Following the batch procedure, accurately weighed quantities of The BET (Brunauer, Emmet and Teller) surface area of the
adsorbent were transferred into 0.8-L bottles, where V = 0.5 L of untreated and the pretreated pine sawdust was measured from the
adsorbate solution were added. The sorbent weight varied from N2 adsorption isotherm with a Nova® Surface Area Analyzer (Quan-
m = 0.5 to 3 g, the temperature was T␪ = 23 ◦ C, the initial MB (MERCK, tachrome Instruments) in accordance with DIN 66132 [30]. Prior to
C.I. 52015) concentration varied from C0 = 1.4 to 156 mg L−1 , the this measurement, the samples were dried under vacuum at 150 ◦ C
initial BB (SIGMA-ALDRICH, C.I. 21000) concentration varied from overnight to clean the surface. The experimental data were ana-
C0 = 2.2 to 11 mg L−1 , and the initial AO (SIGMA-ALDRICH, C.I. lyzed using the “Quantachrome NovaWin2 - Data Acquisition and
46005) concentration varied from C0 = 1.8 to 9 mg L−1 . The bot- Reduction for NOVA instruments” software.
tles were sealed and mechanically tumbled for a period of 7 The scanning electron microscope (SEM) used herein was a JEOL
days. This time period was chosen after experimental studies JSM-6700F Field Emission Scanning Electron Microscopy. The SEM
(the time varied from 4 h to 14 days), to ensure that nearly tests were carried out on samples which were Pt coated (5 nm). The
equilibrium conditions were achieved. The resulting solution con- magnifications were 750, 7500 and 50000, respectively.
centrations were determined and the equilibrium data from each Fourier transform infrared (FTIR) spectra were obtained
bottle represented one point on the adsorption isotherm plots. using a spectroscope (MAGNA-IR 750 Spectrometer, Serrie II,
The initial pH of the solution was 7.8–8.0 and the final pH was Nicolet). The sampling technique used herein was diffuse
7.3–7.7. reflectance. The powder samples were scanned for wavenumber
650–4000 cm−1 .
2.3. Kinetic studies The powder XRD patterns of the origin and the pretreated sam-
ples were measured on by a SIEMENS D5005 X-ray Diffractometer
Adsorption rate batch experiments were conducted in a 2-L using Ni-filtered CuKa ( = 0.154 nm) radiation at 45 kV and 40 mA
completely mixed glass reactor fitted with a twisted blade-type and continuous scan mode. The XRD patterns were recorded in the
stirrer, operating at 300 rpm for keeping the lignocellulosic mate- scan range 2 = 5–70, at scan rate step = 0.04◦ , dwell time = 3 s, i.e.
rial in suspension. The reactor, containing V = 1 L aqueous dye total scan time approximately 1 h and 30 min.
solution, was placed into a water bath to keep temperature constant The concentration of MB, BB and AO in the solution was
at the desired level. The sorbent weight varied from m = 1 to 6 g, the obtained by measuring O.D. at  = 663, 460, and 490 nm, respec-
temperature varied from T␪ = 10 to 55 ◦ C, the initial MB concentra- tively, using a HACH DR4000U UV–vis spectrophotometer. These
tion varied from C0 = 1.4 to 156 mg L−1 , the initial BB concentration optical wavelengths, determined by performing the ‘scan ’ option
varied from C0 = 2.2 to 11 mg L−1 , and the initial AO concentration of the measuring instrument, were satisfactorily compared with
varied from C0 = 1.8 to 9 mg L−1 . The effect of stirring was studied the reported ones in literature as well as with the corresponding
in the rage of 0–600 rpm. The pH effect was studied in the range of values quoted in the dye manufacturer’s manual.
886 D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896

Fig. 2. SEM micrographs for untreated (a–c) and autohydrolyzed (240 ◦ C, 40 min) sawdust (d–f).

3. Results and discussion vs. autohydrolysis reaction time (ta in min, including the preheating
period).
3.1. Autohydrolysis The hemicelluloses content of the untreated pine sawdust was
significantly decreased by autohydrolysis. The cellulose fraction
3.1.1. Autohydrolysis solid residue decreased by its amorphous part (and partially by its crystalline
Cellulose crystallinity, accessible surface area, protection by part) while the lignin quantity remained practically unchanged;
lignin, and sheathing by hemicelluloses, all contribute to lignocel- the lignin percentage on the solid residue increases as is shown in
lulosic biomass resistance to autohydrolysis [1,2]. Autohydrolysis Fig. 1(b).
of cellulose and hemicelluloses produces glucose, manose, xylose
and degradation products as 5-HMF and furfural, respectively. 3.1.2. Microstructure
The autohydrolysis is affected by temperature and reaction The SEM micrographs for untreated pine sawdust and for auto-
time. Cellulose fractions are hydrolyzed to water-soluble cellulo- hydrolyzed at 240 ◦ C for 40 min pine sawdust are given in Fig. 2
oligosaccharides and glucose. Hemicelluloses are hydrolyzed to for three different magnifications. In Fig. 2(a) and (d) with magni-
mano-oligosaccharides, xylo-oligosaccharides, manose and xylose. fication 750 we observe that the pine sawdust particle has fibrous
The acid-insoluble lignin fraction is not significantly affected by structure. This structure is more open for the particle presented in
autohydrolysis. Fig. 2(d), i.e. the autohydrolyzed particle. Comparing Fig. 2(b) and
The yield of the pretreated pine sawdust (dry weight of product (c) with magnification 7500 we observe that the micro fibril struc-
% w/w of the untreated dry material) is presented in Fig. 1, plotted ture of the autohydrolyzed particle is significantly different that
D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896 887

Fig. 3. SEM micrographs for autohydrolyzed (200 ◦ C, 40 min) sawdust before (a–c) and after (d–f) MB adsorption (C0 = 156 mg L−1 , t = 190 min, m/V = 1 g L−1 , 300 rpm, 23 ◦ C).

the untreated one. In Fig. 2(b) we can see the micro fibrils, less than sawdust particle in Fig. 3(a) has some pores. The surface texture of
1 ␮m, but in (d) we can see just a rough texture. For magnification this particle is less rough than that of the autohydrolyzed particle
50,000 we clearly observe that the texture of the pretreated saw- in Fig. 2(f). That fact can be attributed to the less severe autohydrol-
dust particle in Fig. 2(f) is rougher, like to be constructed by smaller ysis conditions. The adsorption conditions were C0 = 156 mg L−1 ,
particles less than 1 nm, comparing to the still smooth surface of the t = 190 min, m/V = 1 g L−1 , agitation speed = 300 rpm. The texture of
untreated one in (f). The picture of the texture in Fig. 2(c) is similar the pretreated sawdust after MB adsorption in Fig. 3(f) is rougher
with that of Klason lignin. This observation seems reasonable, tak- comparing to the same material before MB adsorption presented in
ing into account that the pretreated material is 60.5% (w/w) lignin, Fig. 3(c). This fact indicates the swelling effect on the lignocellulosic
as is presented in Fig. 1(b). These changes can be explained by the material’s particles after MB adsorption.
removal of the hemicelluloses that gave the look to the smooth
particle surface, and the appearance of the naked cellulose-lignin 3.1.3. Surface chemistry
complex which has a more similar to lignin look. The changed sur- The cell walls of pine sawdust mainly consist of cellulose, hemi-
face of the pretreated sawdust explains the enhanced adsorption celluloses, lignin and many hydroxyl groups, such as tannins or
properties and the higher BET surface area values. Moreover, these other phenolic compounds. Lignin is a polymer material built up
particles seem to have cavities but they have not pores. Sawdust is from the phenyl propane nucleus, an aromatic ring with a three-
not a porous material. carbon side chain. Vanillin and syringaldehyde are the two other
In Fig. 3 we can see the SEM micrographs for autohydrolyzed at basic structural units of lignin molecule. Tannins are complex poly-
200 ◦ C for 40 min pine sawdust before and after MB adsorption. The hydric phenols, which are soluble in water and have the property
888 D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896

Fig. 5. (a) XRD patterns of the untreated and the autohydrolyzed (240 ◦ C, 40 min)
sawdust and (b) BET surface area vs. autohydrolysis time.
Fig. 4. FTIR spectra of (a) the untreated and the autohydrolyzed (240 ◦ C, 40 min)
sawdust, and (b) the autohydrolyzed (200 ◦ C, 40 min) sawdust before and after MB
adsorption (C0 = 156 mg L−1 , t = 190 min, m/V = 1 g L−1 , 300 rpm, 23 ◦ C).
indicated the changes of those functional groups on the surface of
autohydrolyzed pine sawdust comparing to the untreated one (see
of precipitating protein (gelatin). The FTIR spectra of the untreated Table 1). These shifts were less severe than those in the case of the
pine sawdust and of the autohydrolyzed (T␪a = 240 ◦ C, tai = 40 min) autohydrolyzed at T␪a = 240 ◦ C pine sawdust for the same time. In
pine sawdust are given in Fig. 4(a). The comparison of the FTIR the case of the autohydrolyzed (T␪a = 200 ◦ C, tai = 40 min) pine saw-
spectrums shows that some peaks were shifted. In the case of dust after MB adsorption, there was a shift at 1454 cm−1 due to OH
the untreated pine sawdust, there is a strong peak at wavenum- bending, at 1088 cm−1 due to C O stretching of six-member cyclic
ber 3471 cm−1 representing the OH stretching of phenol group ether group of cellulose, and at 889 cm−1 indicating the presence
of cellulose and lignin, and the peak at 2939 cm−1 indicates the of aromatic C H deformation (see Table 1). The shifts observed in
presence of CH2 stretching of aliphatic compound. The peak at the FTIR spectrum of the sawdust before and after MB adsorption
2146 cm−1 indicates the presence of NH stretching. The appear- indicated some changes of functional groups on the surface of the
ance of peaks at 1741 cm−1 and 1612 cm−1 indicate the presence sawdust due to bonding with the adsorbed MB.
of C O stretching of aldehyde group and C C stretching of phe-
nol group, respectively. Whereas the peak at 1516 cm−1 in the 3.1.4. Crystalline structure
spectrum of sawdust can be due to C C of aromatic ring. Peaks XRD is applicable to analyze the crystallinity of cellulose in
at 1470 cm−1 and 1437 cm−1 can be due to CH2 bending and wood. XRD patterns of (i) the untreated pine sawdust and (ii)
OH bending, respectively. The peak at 1372 cm−1 shows C O H the autohydrolyzed (240 ◦ C, 40 min) pine sawdust are shown in
bending. Peaks at 1286 cm−1 and 1140 cm−1 in the FTIR spectrum Fig. 5(a). Two broad peaks at the 2 values of around 16◦ and 22◦
of sawdust can be due to C O stretching of phenolic group and for the untreated sawdust were due to the 1 0 1 and 0 0 2 lattice
six-member cyclic ether group of cellulose, respectively. These spacing of cellulose in wood [31]. For the autohydrolyzed sawdust,
wavenumber values are very close to those reported for pine saw- these two peaks became narrow, suggesting structural destruction
dust by Wang et al. [31] and for meranti sawdust by Ahmad et al. of crystalline cellulose, in accordance with a sharp weight loss at
[32]. these conditions (Fig. 1). The obtained rich-to-lignin material is
In the case of the pretreated with autohydrolysis (T␪a = 240 ◦ C, mostly amorphous.
tai = 40 min) pine sawdust, the shifted peaks are: 3359 cm−1 repre-
senting the OH stretching of phenol group, 2065 cm−1 indicating 3.1.5. BET surface area as affected by the autohydrolysis
the presence of NH stretching, 1713 cm−1 indicating the pres- conditions
ence of C O stretching of aldehyde group, 1450 cm−1 due to OH The surface area (in m2 g−1 , measured by the BET method)
bending, 1065 cm−1 due to C O stretching of six member cyclic is given in Fig. 5(b) as a function of autohydrolysis time ta at
ether group of cellulose, and a new one at 822 cm−1 indicating the T␪a = 160–240 ◦ C. The effect of autohydrolysis is not significant at
presence of aromatic C-H deformation (see Table 1). These shifts 160 ◦ C. The BET surface area is increasing significantly for autohy-
observed in the FTIR spectrum indicated the changes of those func- drolysis at 240 ◦ C for ta = 0–50 min.
tional groups on the surface of autohydrolyzed pine sawdust. The efficient removal of the hemicelluloses (in the form of water
In Fig. 4(b) the transmittance of the FTIR spectra of the soluble reducing sugars) and the amorphous part of cellulose (in the
autohydrolyzed (T␪a = 200 ◦ C, tai = 40 min) pine sawdust before form of water soluble glucose) results in ‘opening’ of the pores of
and after MB adsorption are given. The adsorption conditions the structure of the lignocellulosic matrix [9,21]. This phenomenon
were C0 = 156 mg L−1 , t = 190 min, adsorbent = 1 g L−1 , agitation increases the BET surface area of the material, which accounts par-
speed = 300 rpm. In the case of the autohydrolyzed pine sawdust tially for the advanced adsorption properties of the hydrolyzed
before MB adsorption, the shifts observed in the FTIR spectrum materials over the untreated one. Furthermore, the treatment of
D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896 889

Table 1
FTIR of (a) the untreated and the autohydrolyzed (240 ◦ C, 40 min) pine sawdust, and (b) the autohydrolyzed (200 ◦ C, 40 min) pine sawdust before and after MB adsorption.

Frequency (cm−1 ) Differences (cm−1 ) Assignment

Untreated Pretreated

(a) Effect of autohydrolysis at 240 ◦ C for 40 min on sawdust


3471 3359 112 OH stretching of phenol group
2939 2945 −6 CH2 stretching of aliphatic compound
2146 2065 81 NH stretching
1741 1713 28 C O stretching of aldehyde group
1612 1610 2 C C stretching of phenol group
1516 1518 −2 C C of aromatic ring
1470 1468 2 CH2 bending
1437 1450 −13 OH bending
1372 1375 −3 C O H bending
1286 1284 2 C O stretching of phenolic group
1140 1065 75 C O stretching of six-member cyclic ether group
– 822 – C H deformation

Frequency (cm−1 )

Before MB adsorption After MB adsorption

(b) Effect of methylene blue adsorption on autohydrolyzed (200 ◦ C, 40 min) sawdust


3392 3398 −6 OH stretching of phenol group
2916 2916 0 CH2 stretching of aliphatic compound
2156 2162 −6 NH stretching
– – – C O stretching of aldehyde group
1606 1605 1 C C stretching of phenol group
1516 1516 0 C C of aromatic ring
1466 1468 −2 CH2 bending
1433 1454 −21 OH bending
1338 1340 −2 C O H bending
1282 1286 −4 C O stretching of phenolic group
1101 1088 13 C O stretching of six-member cyclic ether group
901 889 12 C H deformation

the material leads to the activation of the internal surface of pine found to be 2.500 for untreated and 2.582 for pretreated sawdust
sawdust, thus increasing the number of active sites available for (see Table 2). The standard error of estimates (SEE)-values were
dye binding. calculated by the following equation

 n
3.2. Adsorption isotherms 
 (yi − yi,theor )2

The comparison of the adsorption capacity of the untreated and  i=1
SEE = (3)
pretreated pine sawdust samples was based on the Freundlich [33], n − p
Langmuir [34] and Sips [35] isotherm models. The first two models
are both widely used for investigating the adsorption of a plethora where yi is the experimental value of the depended variable, yi,theor
of dyes on various lignocellulosic materials and activated carbons. is the theoretical (estimated) value of the depended variable, n
The Freundlich [33] isotherm is given by the following equation: is the number of the experimental measurements and p is the
number of parameters, i.e., (n –p ) is the number of the degrees
q = KF · (Ce )1/n (1) of freedom.
where q is the amount adsorbed per unit mass of the adsorbent The KF for the removal of MB by adsorption on pine sawdust as
(mg g−1 ), Ce is the equilibrium concentration of the adsorbate affected by the autohydrolysis conditions was found in this work
(mg L−1 ) and KF [(mg g−1 )(L mg−1 )1/n ], n are the Freundlich con- to be as follows
stants related to adsorption capacity and intensity, respectively. KF = KF,0 ebKF ta (4)
Deriving the logarithmic form of eq. (1):
1 where ta is the autohydrolysis time measured in min, KF,0 and bKF
log q = log KF + log Ce (2) are empirical parameters; KF,0 = 5.602, coefficients of determina-
n
tion R2 = 0.8467–0.9836 for the corresponding linearized model,
The KF and n values were estimated by non-linear regression and
analysis (NLRA) from the experimental adsorption data obtained
at 23 ◦ C for MB, BB and AO. Fig. 6(a) presents, as an example, MB bKF = 4 × 10−7 Ta 2 − 0.0004Ta + 0.0854 (5)
adsorption isotherms by untreated (in its natural form) and auto-
hydrolyzed (240 ◦ C, 50 min) pine sawdust. The theoretical curves where Ta is the autohydrolysis temperature measured in Kelvin.
are estimated according to the Freundlich equation. The Freundlich The Langmuir isotherm equation [34] is based on the following
parameter values are shown in Table 2. The KF values estimated ‘pseudo-monolayer’ adsorption model.
for the autohydrolyzed samples were significantly higher com- KL qm Ce
paring to those of the untreated material, indicating an increased q= (6)
1 + KL Ce
adsorption capacity of the former; the KF values were exponen-
or
tially increased by increasing the treatment time for constant
1
 1   1
 1
autohydrolysis temperature (see Fig. 7(a)). The parameter n was = + · (7)
not significantly affected by the pretreatment conditions and was q qm KL · qm Ce
890 D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896

Table 2
Estimated values of SEE, when least squares are used, for the alternative isotherm models considered herein.

KF [(mg g−1 )(L mg−1 )1/n ] KL (L mg−1 ) qm (mg g−1 ) n SEE

Methylene blue
Untreated sawdust
Freundlich 5.602 2.500 2.900
Langmuir 0.06842 38.72 1.209
Sips 0.06055 40.42 1.084 1.198
Pretreated sawdust: autohydrolysis at 240 ◦ C for 40 min
Freundlich 15.68 2.582 3.334
Langmuir 0.1072 88.02 4.355
Sips 0.0245 138.1 1.691 1.525
Bismarck brown
Untreated sawdust
Freundlich 3.367 1.683 0.2365
Langmuir 0.3771 12.24 0.2971
Sips 0.02901 40.09 1.482 0.2494
Pretreated sawdust: autohydrolysis at 240 ◦ C for 40 min
Freundlich 15.20 1.797 0.8645
Langmuir 2.567 18.26 1.137
Sips 0.02619 138.2 1.721 0.9299
Acridine orange
Untreated sawdust
Freundlich 3.112 1.802 0.2590
Langmuir 0.5691 8.792 0.2628
Sips 0.02713 33.13 1.595 0.2653
Pretreated sawdust: autohydrolysis at 240 ◦ C for 40 min
Freundlich 15.52 1.435 0.4763
Langmuir 1.914 18.78 0.5210
Sips 0.04112 166.0 1.388 0.5030

where KL is the Langmuir constant related to the energy of adsorp-


tion (L mg−1 ) and qm the amount of dye adsorbed (mg g−1 ) when
saturation is attained. In cases where the isotherm experimental
data approximates the Langmuir equation, the parameters KL and
qm can be estimated either by plotting 1/q versus 1/Ce either by
NLRA. Table 2 presents the estimated (by NLRA) parameter values
for the data gathered in the present study. The qm -values, obtained
for untreated pine sawdust were significantly lower than the values
for the pretreated samples. The fitting of the Langmuir’s adsorption
model to the present data (see theoretical curves in Fig. 6(a)) was
also quite satisfactory but to a lesser degree comparing with the
Freundlich model in the case of the pretreated sawdust, as shown
by the corresponding SEE-values given in Table 2.

Fig. 6. Isotherms of (a) MB, (b) BB and (c) AO adsorption on untreated and autohy-
drolyzed (240 ◦ C, 40 min) sawdust.

Fig. 7. (a) KF and (b) qm for MB adsorption on sawdust vs. autohydrolysis time.
D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896 891

Fig. 10. (a) MB, (b) BB and (c) AO concentration for adsorption on sawdust vs. initial
pH of the dye solution (t = 190 min; C0 = 12.5 mg L−1 for MB, 3.7 mg L−1 for BB and
6.7 mg L−1 for AO; m/V = 1 g L−1 ; 300 rpm; 23 ◦ C).

where Ta is the autohydrolysis temperature measured in Kelvin.


The essential characteristics of the Langmuir isotherm can be
described by a dimensionless constant called ‘equilibrium param-
eter’ or ‘separation factor’ RL , defined by the following equation:

1
RL = (10)
1 + KL · C0

where C0 is the initial dye concentration (mg L−1 ) and KL is the


Langmuir constant (L mg−1 ). The value of RL indicates the type of the
isotherm to be either unfavorable (RL > 1), linear (RL = 1), favorable
(0 < RL < 1) or irreversible (RL = 0). At the present work, the RL values
were found 0.0564–0.9126, i.e. 0 < RL < 1 for all dye concentrations
C0 in the range of 1.4–156 mg L−1 for MB, for all adsorbents (pre-
treated and untreated sawdust) studied. They confirmed that the
pine sawdust is favorable for adsorption of dye under conditions
used in this study.
The Sips (Langmuir–Freundlich) [35] isotherm equation, also
examined in the present work, is based on the following adsorption
model.

qm · (KL · Ce )1/n
q= (11)
Fig. 8. (a) MB, (b) BB and (c) AO adsorption kinetics for untreated and autohy-
1 + (KL · Ce )1/n
drolyzed (240 ◦ C, 40 min) sawdust (C0 = 12.5 mg L−1 for MB, 3.7 mg L−1 for BB and
6.7 mg L−1 for AO; m/V = 1 g L−1 ; 300 rpm, 23 ◦ C).
where KL , qm is the Langmuir constants and n the Freundlich con-
stant.
The qm -values, for the removal of MB by adsorption on pine The parameters of the three isotherm models can be obtained
sawdust as affected by the autohydrolysis conditions (see Fig. 7(b)), by NLRA. Table 2 presents the estimated parameter values for the
were found in this work to be experimental data obtained in the present study. The fitting of the
Sips’ adsorption model to the present data (see theoretical curves
qm = qm,0 ebqm ta (8)
in Fig. 6(a)) was the most satisfactory for MB adsorption, better
where ta is the autohydrolysis time measured in min, qm,0 and than the other two isotherm models, in both cases of the untreated
bqm are empirical parameters; qm,0 = 29.62 mg L−1 , coefficients of and pretreated sawdust, as shown by the corresponding SEE-values
determination R2 = 0.6592–0.8742, and given in Table 2. For BB and AO adsorption the Freundlich model
was the best in all cases (see Fig. 6(b) and (c), and Table 2).
bqm = 8 × 10−5 Ta − 0.0314 (9)

3.3. Kinetics of adsorption

The kinetics of adsorption of MB on several lignocellulosic mate-


rials has been extensively studied using various kinetic equations.
However, information on the kinetics of BB [14] and AO on such
materials is scarce. At first place, we used Lagergren equation [36]:

q − qt = q · e−k·t (12)

where q and qt are the amounts of dye adsorbed per unit mass of
the adsorbent (mg g−1 ) at equilibrium time (t→ ∞) and adsorption
time t, respectively, while k is the pseudo-first order rate constant
for the adsorption process (min−1 ). Moreover, q = (C0 − Ce )V/m and
Fig. 9. k values for (a) MB, (b) BB and (c) AO adsorption on sawdust vs. agitation
speed (C0 = 12.5 mg L−1 for MB, 3.7 mg L−1 for BB and 6.7 mg L−1 for AO; m/V = 1 g L−1 ; qt = (C0 − C)V/m, where C, C0 , Ce are the concentrations of methylene
300 rpm, 23 ◦ C). blue in the bulk solution at time t, 0, and ∞, respectively, while m is
892 D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896

Assuming a -order kinetic model:

dq
= k (q − qt ) (14)
dt
From the differential eq. (14) for  =
/ 1 we obtain:
 1/(1−)
qt = q − q1− + ( − 1) k t (15)

The order of the kinetic model was found to be  = 2.87 ± 1.00


for untreated sawdust and  = 1.87 ± 0.22 for pretreated (autohy-
drolysis, 240 ◦ C, 40 min) sawdust.
For  = 2 we have the pseudo-second order kinetic model intro-
duced [37,38] as follows
 −1 1
qt = q − q−1 + k2 t or qt = q − (16)
(1/q) + k2 t

The diagrams for second order kinetics are presented in Fig. 8(a)
for MB, in Fig. 8(b) for BB and in Fig. 8(c) for AO, for untreated and
pretreated (autohydrolysis, 240 ◦ C, 40 min) sawdust. In the case of
MB, the estimated by NLRA values of the second order rate con-
stant k2 were 0.00091–0.00515 min−1 and the SEE-values were
0.157–0.322. All SEE-values were found low, but higher than the
SEE-values of the first-order kinetic model, indicating the higher
applicability of the first-order kinetic equation to the adsorption
of MB on pine sawdust. In the case of BB, the estimated k2 values
were 0.0055–0.4733 min−1 and the SEE-values were 0.013–0.039.
All SEE-values were found lower than the SEE-values of the first-
order kinetic model, indicating the higher applicability of the
second-order kinetic equation to the adsorption of BB on pine
sawdust. The same was found in the case of AO, where k was
0.0030–0.1594 min−1 and SEE was 0.031–0.111.
In all cases, the k-values of the first-order and second-order
kinetic model were not significantly enhanced by the autohydrol-

Fig. 11. (a) MB, (b) BB and (c) AO amount adsorbed on sawdust vs. t ysis pretreatment of pine sawdust. Despite this, the adsorption
(C0 = 12.5 mg L−1 for MB, 3.7 mg L−1 for BB and 6.7 mg L−1 for AO; m/V = 1 g L−1 ; rate dqt /dt is significantly enhanced due to the (q − qt ) difference
300 rpm; 23 ◦ C).
increase (see Fig. 8).
Agitation is an important parameter in sorption phenomena,
influencing the distribution of the solute in the bulk solution and
the weight of the adsorbent used (g), and V is the solution volume the formation of the external boundary film. The effect of stir-
(mL). Further modification of eq. (16) in logarithmic form gives: ring speed (rpm) on the adsorption rate constant k (min−1 ) of
the untreated material was investigated (see Fig. 9). The kinetics
ln(q − qt ) = ln q − k · t (13) seemed to be affected by the agitation speed for values between
0 and 200 rpm, thus confirming that the influence of external dif-
The plots of ln(q − qt ) vs. t for all dyes adsorbent systems were fusion on the sorption kinetic control plays a significant role. In
found to be linear, indicating the possibility of first order nature contrast, the small effect of agitation in the range of 200–600 rpm
of the adsorption process. MB, BB and AO adsorption kinetics showed that external mass transfer was not the rate limiting step,
by untreated and pretreated (autohydrolysis, 240 ◦ C, 40 min) pine and implied that intraparticle diffusion resistance needed to be
sawdust are presented in Fig. 8(a–c) respectively. The theoret- included in the analysis of overall sorption [21].
ical curves are estimated according to the Lagergren equation. The effect of the pH of the dye solution on the amount of dye
In the case of MB, the estimated by NLRA values of k were adsorbed was studied for MB, BB and AO by varying the initial pH
0.0040–0.0184 min−1 and the SEE-values were 0.100–0.301. All under constant process parameters. The final concentration C of
SEE-values were found low, indicating the applicability of this dyes solutions after an adsorption period of 190 min was signifi-
kinetic equation to the adsorption of MB on pine sawdust. In the cantly higher for pH values between 1.5 and 4 for both untreated
case of BB, the k values were 0.0317–0.1396 min−1 and the SEE- and pretreated materials than the relevant concentration (C for
values were 0.067–0.140, and in the case of AO, the k values were t = 190 min) for pH 8 (see Fig. 10). The lower adsorption of dye
0.0199–0.1594 min−1 and the SEE-values were 0.126–0.233, indi- at acidic pH was due to the presence of excess H+ ions that com-
cating the applicability of this kinetic equation to the adsorption of peted with the dye cation for adsorption sites. As the pH of the
BB and AO on pine sawdust. system increased (pH > 8), the number of positively charged avail-
According to the Lagergren model and the Arrhenius law able sites decreased while the number of the negatively charged
k = p · exp(− E/RT), the activation energy for the adsorption of MB sites increased. The negatively charged sites favored the adsorption
on untreated and pretreated pine sawdust was estimated by lin- of dye cation due to electrostatic attraction. The increase in initial
ear regression of lnk on 1/T. The MB adsorption activation energy E pH from 8 to 13, slightly increased the amount of dye adsorbed.
was found to be 14–15 kJ mol−1 for the treated materials, approx- The final pH of the solution was found to decrease only slightly (by
imately equal to the activation energy of the untreated sawdust 0.3–0.5 pH units) after adsorption of dye (in cationic form) with the
(15 kJ mol−1 ). This indicates that a physical process, i.e. the intra- release of H+ ions from the active site of the adsorbent surface. The
particle diffusion, is the controlling step of the adsorption process. results were in agreement with other literature reports [24–26].
D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896 893

Fig. 12. Breakthrough curves for (a) MB with Ci = 190 mg L−1 , (b) MB with Ci = 14 mg L−1 , (c) BB and (d) AO adsorption on fixed beds (x = 15 cm) of untreated and autohydrolyzed
(240 ◦ C, 40 min) sawdust.

Adsorbate species are probably transported from the bulk of values were found higher than the SEE-values of the first-order
the solution into the solid phase through an intra-particle diffu- and second-order kinetic models, indicating the lower applicabil-
sion/transport process, which is frequently the rate-limiting step ity of the intra-particle diffusion equation to the adsorption of
in many adsorption processes, especially in a rapidly stirred batch MB on pine sawdust. On the other hand, using linear regression
reactor. The possibility of intra-particle diffusion was explored by analysis, qt = 1.2921 + 0.6241t0.5 with coefficient of determination
using the intra-particle diffusion model [39]: R2 = 0.9264 for untreated sawdust, and qt = 2.3447 + 0.946t0.5 with
√ R2 = 0.9071 for pretreated (autohydrolysis, 240 ◦ C, 40 min) sawdust
qt = c + kp · t (17)
(see Fig. 11(a)). Analogous findings were in the case of BB, where
where qt is the amount (mg g−1 ) of dye adsorbed at time kp was 0.057–0.160 mg g−1 min−0.5 and SEE was 0.062–0.179, and
t, c is a constant (mg g−1 ) and kp is the intra-particle diffusion in the case of AO, where kp was 0.342–0.534 mg g−1 min−0.5 and
rate constant in mg g−1 min−0.5 . In the case of MB, the esti- SEE was 0.293–1.317. Moreover, using linear regression analy-
mated by NLRA values of the kp were 0.315–0.805 mg g−1 min−0.5 sis, in the case of BB, qt = −0.0004 + 0.0758t0.5 with R2 = 0.9739 for
and the SEE-values were 0.308–0.596, assuming c = 0. All SEE- untreated sawdust, and qt = 0.1475 + 0.1571t0.5 with R2 = 0.9464 for

Table 3
Estimated values of SEE, when least squares are used, for the alternative column models considered herein.

Ci (mg L−1 ) Q (mL min−1 ) n N (mg L−1 ) K (L mg−1 min−1 ) SEE

Methylene blue
Untreated sawdust
Bohart-Adams 192 46 8046 0.00050 7.58
Bohart-Adams 14.0 10 8722 0.00134 0.194
Clark 192 46 2.500 8723 0.00058 8.75
Clark 14.0 10 2.500 8781 0.00159 0.201
Pretreated sawdust: autohydrolysis at 240 ◦ C for 40 min
Bohart-Adams 187 41 14157 0.00034 7.13
Bohart-Adams 14.3 10 14313 0.00072 0.179
Clark 187 41 2.582 15132 0.00042 8.47
Clark 14.3 10 2.582 14429 0.00087 0.217
Bismarck brown
Untreated sawdust
Bohart-Adams 11.0 10 6576 0.00045 0.300
Clark 11.0 10 1.683 6394 0.00040 0.340
Pretreated sawdust: autohydrolysis at 240 ◦ C for 40 min
Bohart-Adams 11.1 10 8918 0.00093 0.150
Clark 11.1 10 1.797 8866 0.00086 0.156
Acridine orange
Untreated sawdust
Bohart-Adams 8.9 10 4452 0.00168 0.242
Clark 8.9 10 1.802 4425 0.00156 0.262
Pretreated sawdust: autohydrolysis at 240 ◦ C for 40 min
Bohart-Adams 9.0 10 7161 0.00114 0.124
Clark 9.0 10 1.435 6998 0.00093 0.149
894 D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896

Table 4
The Freundlich and Langmuir parameters of adsorption isotherms for various adsorbents according to the literature.

Freundlich Langmuir Reference

KF [(mg g−1 )(L mg−1 )1/n ] n qm (mg g−1 ) KL (L mg−1 )

Methylene blue
Bamboo dust activated carbon 0.68 1.96 143 0.12 [23]
Banana peels 1.34 3.0 20.8 16.5 [16]
Beech sawdust 6.05 1.59 9.78 1.60 [21]
Beech sawdust CaCl2 -treated 11.99 1.61 16.05 2.13 [9]
Beech sawdust MgCl2 -treated 2.91 1.41 14.2 0.288 [12]
Beech sawdust prehydrolyzed 23.42 1.60 30.50 2.09 [21]
Beech sawdust NaCl-treated 3.18 1.61 9.2 0.613 [12]
Beech sawdust ZnCl2 -treated 2.85 1.35 11.7 0.292 [12]
Brazil nut shells 0.216 0.876 7.81 0.15 [42]
Cedar sawdust 92.78 3.94 111.97 20.89 [13]
Cherry 39.84 1.08 [24]
Coconut bunch waste 13.45 3.66 70.92 0.029 [25]
Commercial activated carbon 6.7 3.4 980 0.48 [23]
Crushed brick 44.93 2.83 80.60 1.73 [13]
Hazelnut 76.9 0.36 [24]
Hazelnut shell 21.44 0.65 [26]
Meranti sawdust 12.14 1.441 120.48 0.042 [32]
Oak 29.94 0.30 [24]
Orange peels 1.75 3.8 18.6 19.9 [16]
Palm kernel fiber 8.67 1.77 95.4 0.0317 [17]
Peanut hull dehydrated 81.6 22.3 108.6 0.121 [47]
Pine sawdust 5.602 2.500 38.72 0.06842 In this work
Pine sawdust autohydrolyzed 15.68 2.582 88.02 0.1072 In this work
Pitch-pine 27.78 0.43 [24]
Tea (rejected) 3.15 2.19 147 0.047 [22]
Tea (rejected) NaOH-modified 40.87 2.11 242.11 0.13 [48]
Walnut 59.17 0.15 [24]
Wheat straw 1.59 0.97 2.23 0.758 [11]
Wheat straw prehydrolyzed 11.95 1.33 20.41 0.663 [11]
Wheat straw citric acid-modified 69.14 2.76 396.9 0.0685 [46]
Wood apple shell 20.4 3.23 95.20 0.059 [43]

Bismarck brown
Chemical activated carbon 54 3.07 164 0.22 [14]
Commercial activated carbon 103 4.02 303 0.19 [14]
Fodder yeast cells magnetically modified 8.10 0.394 75.71 0.025 [45]
Grain magnetically modified 7.36 0.418 72.4 0.029 [44]
Pine sawdust 3.367 1.683 12.24 0.3771 In this work
Pine sawdust autohydrolyzed 15.20 1.797 18.26 2.567 In this work
Rubberwood sawdust chemical/steam-activated carbon 598 9.68 1111 0.11 [14]
Rubberwood sawdust steam-activated carbon 1003 8.58 2000 0.07 [14]

Acridine orange
Fodder yeast cells magnetically modified 21.38 0.252 62.22 0.443 [45]
Pine sawdust 3.112 1.802 8.792 0.5691 In this work
Pine sawdust autohydrolyzed 15.52 1.435 18.78 1.914 In this work

pretreated sawdust (see Fig. 11(b)). Similarly, in the case of AO, expressed as non-linearized equation has the form of the ‘simple
qt = −0.1134 + 0.3955t0.5 with R2 = 0.9522 for untreated sawdust, logistic function’:
and qt = 2.0742 + 0.4211t0.5 with R2 = 0.7366 for pretreated sawdust
Ci
(see Fig. 11(c)). In all cases, i.e. MB, BB and AO adsorption on treated C= (19)
1 + Ae−rt
and untreated sawdust, the high R2 values indicate intra-particle
diffusion as the controlling step. where A = eKNx/u ; r = K · Ci .
Clark [41] has developed an alternative to the ‘simple logistic
function’, called the ‘generalized logistic function’, which incorpo-
rates the parameter n of the Freundlich adsorption isotherm:
3.4. Column studies

1/(n−1)
Ci n−1
The ‘bed depth service model’ developed by Bohart and Adams C= (20)
1 + Ae−rt
[40] is as follows
where n = inverse of the slope of the Freundlich isotherm. This
C  K ·N·x expression coincides with Eq. (19) for n = 2. Evidently, we can esti-
i
ln −1 = − K · Ci · t (18) mate the parameters K and N by applying linear and non-linear
C u
regression through expressions (18) and (19), respectively. Since
the SEE is given by expression (3), we may conclude that this
in which C = effluent concentration (mg L−1 ); Ci = influent con- statistic measure of goodness of fitting has a lower value when per-
centration (mg L−1 ); K = adsorption rate coefficient (L mg−1 min−1 ); forming non-linear regression. The equations used to calculate the
N = adsorption capacity coefficient (mg L−1 L); x = bed depth (cm); parameter values after having performed non-linear regression are
u = linear velocity (cm min−1 ); and t = time (min). This model, K = r/Ci and N = u lnA/(xK)= Ci u lnA/(xr).
D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896 895

The values of A and r according to the Bohart–Adams model can of how and to what extent autohydrolysis affects the solid residue
thus be estimated by NLRA from the column effluent data for all the properties. It is a cost-effective method, because the pretreat-
cases presented in Fig. 12(a) and (b) for MB, (c) for BB and (d) for ment expenses are covered by the produced fermentable sugars for
AO. The effluent dye solution volume V (in L) is V = Q. t, where Q is the bioethanol production industry, which is subsidised in the EU
the dye solution flow rate. The theoretical estimations, according Countries. Moreover, the exclusive use of water is cost- and quality-
to the Bohart–Adams model, sufficiently simulate the experimen- effective. The environmental implications might be the relatively
tal data in all the examples given in Fig. 12. The N and K values are high temperatures used (e.g. compared with acid hydrolysis) mak-
calculated from A and r values and presented in Table 3. It can be ing autohydrolysis an energy intensive process. However thermal
mentioned that N = 14157 mg L−1 for the MB adsorption on auto- energy demand can be covered by using cylindrical parabolic con-
hydrolyzed (240 ◦ C, 40 min) sawdust higher than N = 8046 mg L−1 centrators of solar radiation which is adequate for this purpose,
for the untreated one. In addition, K = 0.00034 L mg−1 min−1 for even in the Central-Northern European Countries. Thus, autohy-
the treated material close to the K = 0.00050 L mg−1 min−1 for the drolyzed pine sawdust could be made widely available for use as an
untreated one. The SEE-values are shown in Table 3. According to alternative to commercial activated carbons for the removal of basic
them, the fitting of the Bohart–Adams model to experimental data dyes from water/wastewater effluents. Furthermore, as sawdust is
was found to be better than the fitting of the Clark model for the an industrial waste and no addition of chemicals is required, we
three dyes used herein. According to both models and referring to argue that this process of adsorbent modification may be consid-
the three dyes used, N for the treated pine sawdust was higher than ered to take place within an ‘Industrial Ecology’ framework, since a
that for the untreated one. The incorporated in the Clark model (solid) waste is used to treat another (aquatic) waste, contributing
parameter n of the Freundlich adsorption isotherm, presented in to pollution abatement without entailing excessive cost. This is an
Table 3, was taken from Table 2 results. argument a fortiori, when additives (e.g., chloride salts) are used to
enhance adsorptivity on condition they are available in the vicinity
as waste or low value by-products of proper quality.
3.5. Discussion

The results presented herein showed to which extent the batch Acknowledgments
and column kinetics of MB, BB and AO adsorption on autohy-
drolyzed (160–240 ◦ C, 0–50 min) pine sawdust was improved, The authors acknowledge the financial support from (a) the
comparing to those of the untreated material as control. The Research Center of the University of Piraeus, (b) the Institut fuer
effect of the autohydrolysis on the microstructure and the crys- Kern- und Energietechnik and (c) the University of Minnesota.
tallinity of pine sawdust was significant, as determined by using Technical and economic support has been also provided within the
SEM, FTIR, XRD and BET surface area techniques; more specifi- framework of the independent project “Rational Design of Inno-
cally, the increased and rough surface of the pretreated sawdust vative Catalytic Technologies for Biomass Derivative Utilization”
(shown in SEM micrographs with magnification 50,000) may be running within the University of Minnesota. Moreover, the authors
a main contributor to the enhanced adsorption properties. The would like to thank the anonymous reviewers for their insightful
estimated values of BET surface area, the KF , the qm and the Bohart- suggestions in improving the initial manuscript.
Adams adsorption capacity coefficient N indicate quantitatively
how the specific pretreatment process presented herein enhanced References
the adsorptivity of the corresponding untreated material.
In Table 4 we compare the MB, BB and AO adsorption capacity [1] J.M. Lee, J. Shi, R.A. Venditti, H. Jameel, Autohydrolysis pretreatment of Coastal
KF and qm of pretreated and untreated pine sawdust with other Bermuda grass for increased enzyme hydrolysis, Biores. Techn. 100 (24) (2009)
6434–6441.
adsorbents derived from agricultural or waste materials, according
[2] D. Nabarlatz, A. Ebringerová, D. Montané, Autohydrolysis of agricultural by-
to the available in the literature data [42–48]. products for the production of xylo-oligosaccharides, Carbohyd. Polym. 69 (1)
The sawdust particles mainly consist of cellulose and lignin, and (2007) 20–28.
[3] A. Shukla, Y.-H. Zhang, P. Dubey, J.L. Margrave, The role of straw in the removal
many hydroxyl-rich compounds, such as tannins or other phenolic
of unwanted materials from water, J. Hazard. Mater. B95 (2002) 137–152.
compounds. All these components exhibit properties contribut- [4] G. Crini, Non-conventional low-cost adsorbents for dye removal: A review,
ing to ion exchange, which is a significant stage in several kinds Biores. Technol. 97 (9) (2006) 1061–1085.
of adsorption, favoring also some penetration of the adsorbate [5] V.K. Gupta, Suhas. Application of low-cost adsorbents for dye removal–A
review, J. Env. Manag. 90 (2009) 2313–2342.
(or its product) beyond the surface of the adsorbent. Since ion [6] V.K. Gupta, P.J.M. Carrott, M.M.L. Ribeiro Carrott, Suhas. Low-Cost Adsorbents:
exchange is combined to adsorption, studies in particular kinetics Growing Approach to Wastewater Treatment—a Review, Cr. Rev. Env. Sc. Techn.
and isotherms, provide information on the mechanism of sorp- 39 (2009) 783–842.
[7] A. Demirbas, Agricultural based activated carbons for the removal of dyes from
tion. These mechanisms are, in general, complicated because they aqueous solutions: A review, J. Hazard. Mater. 167 (2009) 1–9.
implicate the presence of different interactions. In addition, a [8] M. Rafatullaha, O. Sulaiman, R. Hashim, A. Ahmad, Adsorption of methylene
wide range of chemical structures, pH, salt concentrations and the blue on low-cost adsorbents: A review, J. Hazard. Mater. 177 (1–3) (2010)
70–80.
presence of ligands often add to the complication. Some of the [9] F.A. Batzias, D.K. Sidiras, Dye adsorption by calcium chloride treated beech
reported interactions include: ion-exchange, complexation, coordi- sawdust in batch and fixed-bed systems, J. Hazard. Mater. B114(1–3) (2004)
nation/chelation, electrostatic interactions, acid–base interactions, 167–174.
[10] P. Nigam, G. Armour, I.M. Banat, D. Singh, R. Marchant, Physical removal of
hydrogen bonding, hydrophobic interactions, physical adsorption
textile dyes from effluents and solid-state fermentation of dye-adsorbed agri-
and precipitation [3,4]. cultural residues, Biores. Technol. 72 (2000) 219–226.
[11] F. Batzias, D. Sidiras, E. Schroeder, C. Weber, Simulation of dye adsorption on
hydrolyzed wheat straw in batch and fixed-bed systems, Chem. Eng. J. 148 (2–3)
4. Conclusions (2009) 459–472.
[12] F.A. Batzias, D.K. Sidiras, Simulation of methylene blue adsorption by salts-
treated beech sawdust in batch and fixed-bed systems, J. Hazard. Mater. 149
Autohydrolysis comparing other pretreatment methods has the (1) (2007) 8–17.
advantage that is using pure and easily recyclable water; no chem- [13] O. Hamdaoui, Batch study of liquid-phase adsorption of methylene blue using
icals (acids, salts, based organic solvents) are needed. The new cedar sawdust and crushed brick, J. Hazard. Mater. 135 (1–3) (2006) 264–273.
[14] B.G.P. Kumar, L.R. Miranda, M. Velan, Adsorption of Bismark Brown dye on acti-
results provide experimental evidence for enhanced adsorption vated carbons prepared from rubberwood sawdust (Hevea brasiliensis) using
properties of the pretreated material. The gain is the understanding different activation methods, J. Hazard. Mater. 126 (1–3) (2005) 63–70.
896 D. Sidiras et al. / Chemical Engineering Journal 171 (2011) 883–896

[15] S.J. Allen, Q. Gan, R. Matthews, P.A. Johnson, Comparison of optimised isotherm [32] A. Ahmad, M. Rafatullah, O. Sulaiman, M.H. Ibrahim, R. Hashim, Scavenging
models for basic dye adsorption by kudzu, Biores. Technol. 88 (2) (2003) behaviour of meranti sawdust in the removal of methylene blue from aqueous
143–152. solution, J. Hazard. Mater. 170 (1) (2009) 357–365.
[16] G. Annadurai, R.-S. Juang, D.-J. Lee, Use of cellulose based wastes for adsorption [33] H.M.F. Freundlich, Über die adsorption in lösungen, Zeitschrift für Physikalische
of dyes from aqueous solutions, J. Hazard. Mater. B92 (2002) 263–274. Chemie 57 (1906) 385–471.
[17] G.O. El-Sayed, Removal of methylene blue and crystal violet from aqueous [34] I. Langmuir, The constitution and fundamental properties of solids and liquids,
solutions by palm kernel fiber, Desalination 272 (2011) 225–232. J. Am. Chem. Soc. 38 (1916) 2221–2295.
[18] T. Robinson, B. Chandran, P. Nigam, The effect of pretreatments of three waste [35] R. Sips, Structure of a catalyst surface, J. Chem. Phys. 16 (1948) 490–495.
residues, wheat straw, corncobs and barley husks on dye adsorption, Biores. [36] S. Lagergren, Zur theorie der sogenannten adsorption gelöster stoffe, Kungliga
Technol. 85 (2002) 119–124. Svenska Vetenskapsakademiens, Handlingar 24 (1898) 1–39.
[19] V.K. Garg, R. Gupta, A.-B. Yadav, R. Kumar, Dye removal from aqueous solutions [37] Y.S. Ho, J.C.Y. Ng, G. McKay, Kinetics of pollutants sorption by biosorbents:
by adsorption on treated sawdust, Biores. Technol. 89 (2003) 121–124. review, Sep. Purif. Methods 29 (2000) 189–232.
[20] X.Y. Shi, B. Xiao, X.Y. Yang, X.P. Zhou, J.F. Li, Batch study of dye removal from [38] F.-C. Wu, R.-L. Tseng, S.-C. Huang, R.-S. Juang, Characteristics of pseudo-second-
aqueous solutions by adsorption on NaOH-treated firry sawdust, Fres. Env. Bull. order kinetic model for liquid-phase adsorption: A mini-review, Chem. Eng. J.
16 (12A) (2007) 1583–1587. 151 (1–3) (2009) 1–9.
[21] F.A. Batzias, D.K. Sidiras, Dye adsorption by prehydrolysed beech sawdust in [39] W.J. Weber, J.C. Morris, Kinetics of adsorption on carbon from solution, J. Sanit.
batch and fixed-bed systems, Biores. Technol. 98 (6) (2007) 1208–1217. Eng. Div. Am. Soc. Civ. Eng. 89 (1963) 31–60.
[22] N. Nasuha, B.H. Hameed, T. Azam, Din. Mohd, Rejected tea as a potential low- [40] G. Bohart, E.N. Adams, Some aspects of the behavior of charcoal with respect
cost adsorbent for the removal of methylene blue, J. Hazard. Mater. 175 (2010) to chlorine, J. Am. Chem. Soc. 42 (1920) 523–544.
126–132. [41] R.M. Clark, Modeling TOC removal by GAC: The general logistic function, J. Am.
[23] N. Kannan, M.M. Sundaram, Kinetics and mechanism of removal of methylene Wat. Works Assoc. 79 (1) (1987) 33–37.
blue by adsorption on various carbons–a comparative study, Dyes Pigments 51 [42] S.M.O. Brito, H.M.C. Andrade, L.F. Soares, R.P. Azevedo, Brazil nut shells as a
(2001) 25–40. new biosorbent to remove methylene blue and indigo carmine from aqueous
[24] F. Ferrero, Dye removal by low cost adsorbents: Hazelnut shells in comparison solutions, J. Hazard. Mater. 174 (2010) 84–92.
with wood sawdust, J. Hazard. Mater. 142 (1–2) (2007) 144–152. [43] S. Jain, R.V. Jayaram, Removal of basic dyes from aqueous solution by low-
[25] B.H. Hameed, D.K. Mahmoud, A.L. Ahmad, Equilibrium modeling and kinetic cost adsorbent: Wood apple shell (Feronia acidissima), Desalination 250 (2010)
studies on the adsorption of basic dye by a low-cost adsorbent: Coconut (Cocos 921–927.
nucifera) bunch waste, J. Hazard. Mater. 158 (1) (2008) 65–72. [44] I. Safarik, K. Horska, M. Safarikova, Magnetically modified spent grain for dye
[26] M. Dogan, H. Abak, M. Alkan, Biosorption of methylene blue from aqueous solu- removal, J. Cereal Sci. 53 (2011) 78–80.
tions by hazelnut shells: Equilibrium, parameters and isotherms, Water Air Soil [45] I. Safarik, L.F.T. Rego, M. Borovska, E. Mosiniewicz-Szablewska, F. Weyda,
Pollut. 192 (1–4) (2008) 141–153. M. Safarikova, New magnetically responsive yeast-based biosorbent for the
[27] L. Segal, J.J. Greely, A.E. Martin, C.M. Conrad, An empirical method for estimating efficient removal of water-soluble dyes, Enzyme Microb. Tech. 40 (2007)
the degree of crystallinity of native cellulose using the x-ray diffractometer, 1551–1556.
Textile Res. J. 29 (1959) 786–795. [46] R. Han, L. Zhang, C. Song, M. Zhang, H. Zhu, L. Zhang, Characterization
[28] J.F. Saeman, J.F. Bubl, E.E. Harris, Quantitative saccharification of wood and of modified wheat straw, kinetic and equilibrium study about copper ion
cellulose, Ind. Eng. Chem. Anal. Ed. 17 (1945) 35–37. and methylene blue adsorption in batch mode, Carbohyd. Polym. 79 (2010)
[29] Tappi Standards, Tappi Tests Methods, T222 om-88, Atlanta (1997). 1140–1149.
[30] DIN 66132, Determination of specific surface area of solids by adsorption of [47] D. Ozer, G. Dursun, A. Ozer, Methylene blue adsorption from aqueous
nitrogen; single-point differential method according to Haul and Dümbgen solution by dehydrated peanut hull, J. Hazard. Mater. 144 (2007) 171–
(1975). 179.
[31] Z. Wang, J. Cao, J. Wang, Pyrolytic characteristics of pine wood in a slowly [48] N. Nasuha, B.H. Hameed, Adsorption of methylene blue from aqueous solu-
heating and gas sweeping fixed-bed reactor, J. Anal. Appl. Pyrol. 84 (2) (2009) tion onto NaOH-modified rejected tea, Chem. Eng. J. 166 (2011) 783–
179–184. 786.

Вам также может понравиться