Video 21
Video 21
Video 21
PII: S1385-8947(12)01637-3
DOI: http://dx.doi.org/10.1016/j.cej.2012.12.008
Reference: CEJ 10142
Please cite this article as: N.E. Travlou, G.Z. Kyzas, N.K. Lazaridis, E.A. Deliyanni, Graphite oxide/chitosan
composite for reactive dye removal, Chemical Engineering Journal (2012), doi: http://dx.doi.org/10.1016/j.cej.
2012.12.008
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Graphite oxide/chitosan composite for reactive dye removal
∗
Corresponding author. Tel.: +30 2310 997808; fax: +30 2310 997859. E–mail address:
1
ABSTRACT
In the current work a novel composite material consisted of cross–linked chitosan (Ch) and graphite oxide
(GO) (and not graphene oxide, as numerous studies deal with) was prepared for the removal of Reactive
Black 5 dye. After synthesis, the full characterization with various techniques (SEM/EDAX, FTIR, XRD,
DTA, DTG, TGA) was achieved revealing many possible interactions/forces of dye–composite system.
Through the latter interactions, the possible adsorption mechanism was elucidated and explained. Also, the
adsorption evaluation of the composite material presented high adsorption capacity (277 mg/g at 25 oC).
Equilibrium experiments are also performed studying the effect of pH on adsorption (optimum 2) and
desorption (optimum 12), initial dye concentration, contact time (pseudo–first, –second order equations and
a generalized fractal kinetic model), temperature (isotherms at 25, 45 and 65 oC fitted to Langmuir,
Freundlich, Langmuir–Freundlich model) and ionic strength. A full thermodynamic evaluation was carried
out, calculating the parameters of enthalpy, free energy and entropy ( H0, G0, and S0). The adsorption
behavior of the composite material was also compared with the respective of its components (GO and Ch).
Interactions
2
1. Introduction
Nowadays, one of the most promising adsorption methods for the removal of dyes is the application
of chitosan. Chitosan is synthesized from chitin, which is the second most abundant polymeric material in
nature after cellulose [1]. Chitin can be easily extracted from crustacean shell such as crabs, fungi, prawns,
insects and other crustaceans [1]. Chitosan can be used as an adsorbent to remove cationic and/or anionic
dyes due to the simultaneous presence of amino and hydroxyl groups, which can serve as active/adsorption
sites [2]. The performance of chitosan as adsorbent can be improved by (i) the use of cross–linking
reagents, which stabilize chitosan in acid solutions and enhance its mechanical properties and (ii)
derivatization with grafting functional groups onto the chitosan backbone, which improve its adsorption
capacity [3].
Recently, chitosan composites have been developed to adsorb dyes from wastewaters [4,5].
Different kinds of substances have been used to form composite with chitosan such as graphene oxide and
magnetite. These composites have been affirmed to present better adsorption capacity and resistance to
acidic environment. Some works in literature dealt with highlights of the application of chitosan composites
as adsorbents, including synthesis, mechanisms and other factors, which can affect its adsorption capacity
[4,5].
Graphite–like nanoplatelets have recently attracted attention as viable and inexpensive filler in
composite materials. These excellent properties may be relevant at the nanoscale, if graphite can be
exfoliated into thin nanoplatelets, and even down to the single graphene sheet level [6]. Graphene and
graphene oxide have been used as effective adsorbent toward anionic and cationic dyes [7].
Graphite oxide (GO), a precursor in the formation of graphene layers, has been studied as an
adsorbent and its ability to remove cationic dyes has been demonstrated [8,9]. From a structural point of
view, GO is a highly–oxidized planar material containing 25–33% oxygen with good adsorption between
layers. It is characterized to be a lamellar solid and can be easily prepared by chemical modification.
Generally, GO can be synthesized by oxidation of graphite with strong oxidizing agents such as potassium
3
chlorate, permaganate, bichromate and chlorine dioxides. The strong oxidation makes GO to have multiple
oxygen–containing functional groups (as for example carboxyl, hydroxyl, epoxy) and could consequently
be covalently attached to its layers. Moreover, the spatial distribution of functional groups on the carbon
skeleton is also diverse. The epoxy and hydroxyl groups are near each other, and the carboxyl is possibly
situated at the edges of GO. These oxygen functionalities make GO extensively disperse in water, because
The novelty of the current work is the synthesis of the new composite material graphite oxide –
chitosan (GO–Ch), consisted of cross–linked chitosan (Ch) and graphite oxide (GO). The composite
prepared was applied for the removal of the Reactive Black 5, given the lack of literature of a GO/chitosan
material as reactive dye adsorbent. Reactive Black 5 dye has been used in the textile industries for the
dying of cotton, woollen and nylon fabrics worldwide. It is reported to be toxic and cause allergic reactions
of respiratory tract [11]. After synthesis, the characterization with various techniques (SEM/EDAX, FTIR,
XRD, DTA, TGA) was performed revealing many possible interactions of dye–composite system. Through
the latter interactions, the possible adsorption mechanism was elucidated and explained. The influence of
solution pH, initial dye concentration, contact time, temperature and ionic strength was performed in batch
mode.
2.1. Materials
Chitosan of high molecular weight was purchased by Sigma–Aldrich and purified by extraction with
acetone in a Soxhlet apparatus for 24 h. Then, its drying was carried out under vacuum at 20 oC. Its average
molecular weight was estimated at 3.55×105 g/mol and the degree of deacetylation was 82 wt% [12]. The
cross–linking agent used was glutaraldehyde (GLA), which was obtained from Sigma–Aldrich (50 wt% in
A commercial reactive dye (anionic and anthraquinonic) was used as target molecule for dye
4
adsorption experiments. The reactive dye, C.I. Reactive Black 5 – 20505 (abbreviated hereafter as RB5,
λmax=603 nm, purity=55% w/w. The dye purity was taken into account for all calculations. The chemical
Three adsorbent materials were prepared: (i) cross–linked chitosan powder (Ch); (ii) graphite oxide
Ch was prepared in the same manner as in a previously published work [13]. Briefly, 5 g of chitosan
powder was dissolved into acetic acid solution (2% v/v) with the use of a magnetic stirrer suspended in an
0.05 mol/L aqueous solution of cross–linker GLA (approximately 2:1 GLA aldehyde groups (–CHO) per
initial chitosan amine groups (–NH2)). The reactions were carried out at ambient temperature for 24 h and
the cross–linking mechanism is presented in Supplementary Data (Fig. SI1). Extensive rinsing of the
granulated chitosan particles with hot deionized water ensured the removal of any unreacted GLA reagent.
GO was similarly prepared according to a simplified Hummers method [14], following which
commercial graphite powder (10 g) was stirred in concentrated sulfuric acid (230 mL, 0 oC). Then,
potassium permanganate (30 g) was slowly added to the suspension. The rate of addition was controlled to
prevent the rapid rise in the temperature of the suspension that should be less than 20 oC. The reaction
mixture was then cooled to 2 °C. After removal of the ice–bath, the mixture was stirred at room
temperature for 30 min. Distilled water (230 mL) was slowly added to the reaction vessel, keeping the
temperature less than 98 oC. The diluted suspension was stirred for an additional 15 min, was further
diluted with 1.4 L of distilled water and then 100 mL of a 30 wt% solution of hydrogen peroxide was
added. The mixture was left overnight. The GO particles settled at the bottom were separated from the
excess liquid by decantation. The remaining suspension was transferred to dialysis tubes (Sigma Co.).
Dialysis was carried out until no precipitate of BaSO4 was detected by addition of an aqueous solution of
5
BaCl2. Then, the wet form of graphite oxide was separated by centrifugation. The gel–like material was
freeze dried and a fine dark brown powder of the initial graphite oxide was obtained.
For the preparation of the novel GO–Ch composite [8,9], chitosan solution (2% w/v) was prepared
by dissolving 0.4 g of powder chitosan into 20 mL of acetic acid solution (2% v/v) under ultrasonic stirring
for 2 h at room temperature. Also, 3 mL of GLA (50 wt% in water) were added to cross–linked chitosan.
Then, 0.3 g of GO were added in the solution prepared and the mixed system was stirred continuously for
90 min in a water bath at 50 oC. The pH of the reaction system was adjusted to 9–10 with micro–additions
of NaOH (0.1 mol/L) and kept in the water bath for further 60 min at 80 oC. Black products were washed
with ethanol and distilled water in turn until the pH was reached about 7 and dried in a vacuum oven at 50
o
C. The final product was the composite of graphite oxide/cross–linked chitosan (GO–Ch). All prepared
products (GO, Ch, and GO–Ch) were ground to fine powders, with size after sieving 75–125 µm.
Scanning electron microscopy (SEM) images were performed at Zeiss Supra 55 VP. The
accelerating voltage was 15.00 kV and the scanning was performed in situ on a sample powder. EDAX
analysis was done at magnification 10 K and led to the maps of elements and elemental analysis. The FTIR
spectra of the samples were taken with a PerkinElmer–2000 FTIR spectrometer using KBr disks prepared
by mixing 0.5% of finely ground carbon sample in KBr. Pellet made of pure KBr was used as the reference
sample for background measurements. The spectra were recorded from 4000 to 400 cm–1 at a resolution of
4 cm–1. The spectra presented are baseline corrected and converted to the transmittance mode. Also,
thermal analysis was carried out using a TA Instrument thermal analyzer (SDT). The instrument had the
following settings: (i) heating rate of 10 K/min, and (ii) flow rate of nitrogen atmosphere equal to 100
mL/min. Approximately 25 mg of sample was used for each measurement. Moreover, X–ray powder
diffraction (XRD) patterns were recorded on a Philips PW1820 diffractometer with a Cu Kα radiation for
6
2.4. Adsorption–desorption experiments
For the pH–effect experiments, 0.02 g of adsorbent was mixed with 20 mL of dye solution (the
initial dye concentration was 250 mg/L). The suspensions were shaken for 24 h (pH=2–12; ionic
strength=0.001 mol/L; agitation rate=160 rpm) into a water bath to control the temperature at 25 °C (Julabo
SW–21C). The optimum pH found (for performing the next adsorption experiments) was pH=2 for RB5.
To study the effect of contact time on dye adsorption, kinetic experiments were performed by
mixing 0.02 g of adsorbent with 20 mL of dye solution (the initial dye concentration was 250 mg/L). The
suspensions were shaken for 24 h at the optimum pH (found from experiments for the effect of adsorption
pH) in water bath at 25 °C (ionic strength=0.001 mol/L; agitation rate=160 rpm). Samples were collected at
fixed intervals (5–45 min, 1–24 h). Pseudo–first [15], –second order equations [16] and Brouers–Sotolongo
model (BS) [17,18] were selected to fit the experimental kinetic data. The BS model generalized fractal
kinetic equation was developed in order to come with a universal function for the kinetics of complex
systems characterized by stretched exponential and/or power law behaviors. This kinetic function unifies
and generalizes previous theoretical attempts to describe what has been called “fractal kinetic”. The kinetic
The effect of initial dye concentration on equilibrium was realized by mixing 0.02 g of adsorbents
with 20 mL of dye solutions with varying initial dye concentrations (0–1000 mg/L). The suspensions were
shaken for 24 h at pH=2 in water bath at 25, 45, and 65 °C (ionic strength=0.001 mol/L; agitation rate=160
rpm). The equilibrium data resulted were fitted to the Langmuir [19], Freundlich [20] and Langmuir–
Freundlich (L–F) isotherm models [21]. The amount of total dye uptake at equilibrium Qe (mg/g) was
(C0 − Ce )V
Qe = (1)
m
where C0, Ce (mg/L) are the initial and equilibrium concentrations of dye in the aqueous solution; m (g) is
performing experiments with various values of ionic strength (0.001, 0.1, and 1 mol/L NaCl). The other
parameters of adsorption were kept constant (pH=2; 0.02 g of adsorbent; 20 mL of solution volume; 250
Completing adsorption experiments with 250 mg/L of dye solution at 25 °C and pH=2, the dye–
loaded adsorbent material was collected and filtered, using fixed pore–sized membranes (0.50 µm).
m). A small
fraction of the dye (1–2 %) and adsorbent (1%) were retained on the filter membrane; these small variations
due to filtration were neglected. To determine the optimum desorption pH value of the dye–loaded
sorbents, experiments were carried out by mixing the collected, after adsorption, amount of dye–loaded
materials (0.02 g) with 20 mL (same volume as in the adsorption step) of deionized water as eluant over a
pH range between 2 and 12 (pH–adjustment with addition of acid (HNO3) or base (NaOH)), at 25 °C for 24
All experiments were repeated 4 times. Therefore, the experimental points presented in adsorption
The residual dye, i.e. that remaining in solution after the application of solid/liquid separation of
suspended solids using 50 μm pore size filtration membrane, was analyzed spectrophotometrically by
monitoring the absorbance of the dyes using UV–Vis spectrophotometer (model U–2000, Hitachi). Before
the adsorption experiments, the effect of pH over the calibration curves of dye was studied, but no
The XRD patterns of Ch, GO, and GO–Ch composite are presented in Fig. 2, while the X–ray
8
diffraction patterns of pristine graphite and GO (for comparison) are shown in the inset of Fig. 2. Observing
the patterns of inset, the d002 peaks reflect the changes in the nature of graphite by introduction of chemical
groups to the graphene layers. The d002=3.36 Å diffraction peak of natural graphite occurs near the location
where 2θ=26o; it can be seen from the pattern that the peak is sharp and high, showing high crystallinity
and degree of order. In the spectrogram of GO, the diffraction peak (002) of graphite crystal disappears,
indicating that the graphite had been completely oxidized. In its place, the diffraction peak (002) of GO
occurs. The characteristic XRD peak of GO appeared at 2θ=10.9o [22]. It can be seen that the interlayer
distance between the carbon layers as determined by Bragg’s law, is increased clearly from 3.36 Å for
graphite to 8.11 Å for graphite oxide. This is caused by the large amount of polar groups generated between
the layers of graphite during oxidation [23,24], in which the oxygen and carbon atoms are covalently
bonded, leading to an increase in the graphite’s crystal lattice length along axis c. Also, the existence of the
polar groups causes very strong hydroscopicity of GO, as well. As a result, the water molecules entering
into the space between the layers are bonded with the oxygen atoms of the groups through hydrogen bonds,
further enlarging the interlayer distance [25]. On the other hand, the elemental analysis of GO evaluated
using energy dispersive X–ray analysis (EDX), showed that GO had a composition of 48% carbon and
49.5% oxygen (w/w). In addition, 1.5% (w/w) of sulfur is also presented, which is a consequence of
sulfuric acid treatment. Sulfones are also introduced to graphene sheets or residual sulfuric acid is
intercalated between layers. This may also result in defects in the graphene sheets, reflected in the greater
The XRD patterns of Ch, GO, and GO–Ch composite are presented in Fig. 2. Both the patterns of
Ch and GO–Ch exhibited a broad peak at 2θ=20.2o due to the amorphous state of chitosan, indicating that
the amorphous–like structure of Ch was not changed by the addition of GO in the composite material.
Moreover, the feature of GO is still visible at 2θ=10.9o in the XRD pattern of the composite, which
indicates the presence of the two phases in this material. Considering similarly the Bragg’s law, the
interlayer distance of the composite’s component was found to remain 8.11 Å. However, according to
9
Scherrer’s equation, the number of layers were found to decrease from 7.5 for GO to 3.7 for GO–Ch. That
decrease may be the apparent result of the decrease of the average size of the crystallite of the GO–Ch
composite, from 61 Å for the GO to 29 Å for the GO–Ch composite, calculated using the Debye–Scherrer
equation [26].
The insertion pathway of aliphatic amines via the nucleophilic substitution reactions on the epoxy
groups of GO induces changes in the basal spacing of GO respective to their chain length [27]. On the other
hand, the introduction of large organic ammonium ions also resulted in changes in the basal spacing or/and
exfoliation of layered structure [28]. Taking these into account, as well as the XRD results it is concluded
that a significant part of the graphite oxide did not exfoliated/reduced on reacting with chitosan having as
Changes are also obvious in the SEM images of GO and GO–Ch and presented in Fig. 3. The GO
presents the sheet–like structure (Fig. 3a), while after the combination with Ch a rougher surface is
presented (Fig. 3b), revealing that Ch had been assembled on the surface of GO layers with a high density
[8].
Differences in the chemistry of the surfaces have been also seen on differential thermal gravimetric
(DTG) curves measured in nitrogen (Fig. 4). The peaks represent weight loss at the specific temperature
range and the area under the peaks is related to the extent of that weight loss. The first peak centered at
about 80–100 oC for all under examination samples, in correlation to the endothermic effect on the
differential thermal analysis (DTA) curves presented in Fig. 5 can be linked to weight loss due to the
evaporation of physically adsorbed water. In the case of the GO–Ch composite, that peak presented a
maxima at 100 oC, while the second peak presented between 200–250 oC is related to decomposition of
epoxy and carboxylic groups, with latter being more stable [23], that are associated with the exfoliation of
graphene layers, which is represented by exothermic peaks on the differential thermal analysis (DTA) curve
presented in Fig. 5. The shoulder observed between 250–300 oC for the GO represents decomposition of
10
The TGA of Ch, GO and GO–Ch were conducted in nitrogen atmosphere and the curves are shown
in the inset of Fig. 4. GO shows ~20% mass loss below 200 °C, resulting from the evaporation of adsorbed
water and a rapid 26% mass loss from 200 to 250 °C owing to the removal of the oxygen–containing
functional groups. It then displays a gradual mass loss above 250 °C, which is due to the further removal of
functional groups. The mass loss of GO–Ch composite is obviously lower than that of GO, especially the
mass loss at 200 °C (~7%), which demonstrates a gradual decrease in the amount of oxygen–containing
functional groups. Meanwhile, higher mass losses are found in GO–Ch at 400 °C caused by the
decomposition of Ch [29]; this indicates that Ch has been adsorbed onto the GO nanosheets.
As seen from the graphs, there is a significant difference in thermal behavior between the GO–Ch
composite and the separate GO and Ch samples. On the DTG curve of this sample, two peaks with maxima
at 150 and 250 °C are presented. Because the first peak is associated with an exotherm on DTA curves (Fig.
5), it is related to the decomposition of surface functional groups, accompanied by some degree of
exfoliation of the graphene layers [30]. This decomposition occurs at lower temperature than that in the
case of pure GO. The second peak, which is linked to an exotherm, represents the degradation of chitosan
[31]. The fact that this peak appears at higher temperature than for the pure chitosan (Ch) suggested an
FTIR spectra (Fig. 6) revealed the obvious differences in the chemical features of the raw materials,
as well as for the composite one. In the spectrum of GO (spectrum A), carboxyl groups are observed as
bands at 1050–1100 and 1665–1760 cm–1, whereas the band at ~1600 cm–1 is attributed to the presence of
epoxy groups [23] and to C=C stretching mode of the sp2 carbon skeletal network. The peak at 1380 cm–1,
as well as the broad band between 2200 and 3800 cm–1 correspond to O–H groups (C–OH stretching),
which are either attached to carbons or represent adsorbed water. The band at 1055 cm–1 can be also
attributed to epoxy groups. On the spectrum a vibration at 1220 cm–1 is visible, which may be arise from
S=O asymmetric stretching vibrations in sulfones or sulfates present after graphite oxidation and/or
vibrations of C–O in epoxy groups [23,24]. The pattern of cross–linked chitosan (spectrum C) has a strong
11
peak around 3400 cm–1 due to the stretching vibration of O–H, the extension vibration of N–H and the
inter–hydrogen bonds of the polysaccharide [33], while two characteristic bands centered at 1636 and 1597
cm–1 can be attributed to the C=O stretching vibration of NHCO (amide I) and the N–H bending of NH2,
respectively [13]. On the spectrum of GO–Ch composite (spectrum E), the intensity of the peaks at 3417
(νO–H) and 1075 (νC–O) cm–1 increased compared to GO (spectrum A), which is due to the –OH groups from
the adsorbed chitosan. The peak at 1730 cm–1 (attributed to carboxylic groups) disappeared, proving that
carboxylic groups reacted with chitosan during the preparation of the composite, while it can be distinctly
observed that the intensity of NH2 absorbance band has increased compared to the intensity of the amino
group NHCO (amide I). Since chitosan has –NH2 groups on its macromolecular chains, these functional
groups can form hydrogen bonds and/or electrostatic interactions with the residual oxygen–containing
groups of GO. The amino groups of chitosan can react with oxygen–containing functional groups of GO in
the following ways: (i) after hydrogen–bonding interactions between the amino groups and the oxygen–
containing functional groups of GO (G–OH…H2N–R) (Eq. (2a)); (ii) after protonation of the amine by the
weakly acidic sites of the GO layers (–COO–+H3N–R) with amides and amine carboxylate salts as reactions
products (Eq. (2b)), and (iii) after nucleophilic substitution reactions on the epoxy groups of GO (Eq. (2c))
[27]:
H O H O
NH + C RGO N C + H2O
RCh HO RCh RGO (2a)
H O H O
NH + C RGO NH2+O C
RCh HO RCh RGO (2b)
H H
O H+ O RChNH2-H RChNH2 C C OH
RChNH2 C C OH
(2c)
where RCh is the residual part of Ch (especially its amino groups), and RGO is the respective part of GO
12
Concerning the spectra of the GO–Ch composite (spectrum E), it is concluded that carboxyl groups
of GO reacted with –NH2 groups of chitosan to form amides (band at 1636 cm–1) and carboxyl groups
(band at ~1400 cm–1), in line with the amine nucleophilic attack on the epoxy groups of GO, which leads to
the amino (band increase at 1597 cm–1) and hydroxyl (band at 1027 cm–1) formation.
In Fig. 6, the spectra of adsorbents after dye adsorption are presented. For GO, the shift of the band
(spectrum B) is indicative that the binding of dye molecules is predominant based on the π–π interaction
between the aromatic ring of the dye and the GO basal planes. For Ch and GO–Ch composite as well, the
driving forces for reactive dye adsorption could be given by the following equations [34]:
⎯⎯
→ R − NH 3+
R − NH 2 + H + ←⎯
⎯ (3)
⎯⎯
→ R − NH 3+ − O3S − Dye
R − NH 3+ + Dye − SO3− ←⎯
⎯ (5)
The aforementioned interactions between dye molecules and (i) Ch, (ii) GO–Ch, was affirmed by the
diminishing of the amine peak at 1590 cm–1 after comparison of the spectra for initial and after dye
adsorption samples (spectrum D; spectrum F) confirming the binding of the dye molecule to amino groups
according to Eq. (5). The interaction between RB5 and Ch was also affirmed by the shift of the carbonyl
peak. In particular, the amide–carbonyl peak of Ch was shifted from 1659 to 1652 cm–1. For the GO–Ch
composite, the contribution of GO to dye adsorption can be attributed to the π–π dispersion interaction
between the aromatic ring of the dye and the GO basal planes. Due to the binding of carboxyl groups to
Fig. 7 shows the effect of initial pH on the adsorption of RB5 onto adsorbents, which is nearly
explained by the interactions described above. The optimum pH found for all adsorbents tested was
13
acidic (pH=2). This fact is explained by Eq. (5), where the main interaction is the electrostatic force
between amino groups of Ch or Ch–GO and dye. At pH=2, the removal of RB5 was 86, 75, 54% for GO–
Ch, Ch, and GO, respectively. Increasing the pH of the solution, it is absolutely normal the weakening of
forces and consequently the reduction of attraction/interaction between dye molecules and adsorbent. So, at
alkaline conditions (pH=12), the dye removal was 36, 11, 32% for GO–Ch, Ch, and GO, respectively. GO
has not been affected so much from the change of pH (from 54 at pH=2 to 32% at pH=12), which attributed
to the existence of π–π interactions as the main forces between GO and RB5, and not the electrostatic bonds
between positively charged amino groups and negatively sulfonate groups of dye. In general, the pH–
behavior observed was the expectable based on considerations of 3.1 section. Fig. 7 also shows the changes
3.3. Kinetics
The effect of contact time on adsorption is depicted in Fig. 8. The kinetic curves of Ch and GO
obtained were qualitatively same as those in many papers [35]. The instantaneous dye adsorption (0–1 h),
governed by fast external diffusion and mainly surface adsorption, was followed by a milder and gradual
ascend (1–3 h), resulting in an equilibrium state plateau from 3 to 24 h. However, the kinetic behavior of
GO–Ch composite is slightly different. The kinetic rate seemed to be slower than that of GO and Ch. The
first stages lasted approximately 3 h, while the gradual equilibrium reached 15 h. The equilibrium is
observed to be completed after 18 h. This different behavior may be due to the co–existence on the surface
of GO–Ch different active adsorption sites (amino or carboxyl groups), which maybe act competitively.
Also, a possibility of pore diffusion can be posed, given the complicated form (network) of GO–Ch
composite.
Three kinetic models, pseudo–first, –second and Brouers–Sotolongo fractal kinetic, were selected to
fit the experimental kinetic data. The identification of the kinetic model, which best fitted the results has
been performed by non–linear regression and employing experimental Qe values determined independently
(equilibrium data). Table 1 presents the kinetic parameters resulted by fitting the models to the
14
experimental data. According to the correlation coefficients (R2) exported, the best fitting was for the
Brouers–Sotolongo fractal kinetic model (R2~0.999). This could be attributed to the additional parameters n
and a, which serve as extra degrees of freedom to make the BS model more flexible.
Fig. 9 presents the effect of initial dye concentration on adsorption of RB5. The experimental results
were fitted to Langmuir, Freundlich and L–F models. Table 2 reports the maximum adsorption capacities
(Qmax) and the other isothermal parameters resulted from the fitting. The correlation coefficients (R2>0.990)
showed that the L–F model results in better fitting (closer prediction of the isotherm to the experimental
data). The Qmax at 25 oC was 205, 224, and 277 mg/g (pH=2) for GO, Ch, and GO–Ch, respectively.
Furthermore, the dye uptake was influenced by the initial dye concentration (with constant adsorbent’s
dosage). The equilibrium is intense and continues rapidly at low initial dye concentrations. So, there is
possibility for monolayer coverage of dye molecules at the outer interface of adsorbents. Increasing the
initial dye concentrations, the available adsorption sites become lower and subsequently the adsorption
depends on the initial concentration of dye. In the case of GO–Ch, the latter phenomenon is more intense,
The effect of temperature on equilibrium is presented in Table 2. Increasing the temperature from
25 to 65 oC, an increase of the Qmax is observed. As the majority of the chitosan–based and graphite–like
materials [6,7,22,29,35], this increase is caused by the enhancement of the number of adsorption sites; the
latter is due to the weakening and/or breaking of many structural bonds of network, existed near the edge of
the active surface sites of materials. Another, possible explanation is the increase of diffusion (especially
that of pores), which helps the penetration of dye molecules inside the structure of materials. Also, some
works revealed the role of the formation of new active adsorption sites at high temperatures [35]. In
addition, the parameter RL (dimensionless), which is called equilibrium parameter, is calculated according
to Eq. (6) [36] in order to identify whether an adsorption system is favorable or unfavorable.
15
1
RL = (6)
1+K L C0
The range of RL values (Table 2) was calculated for all initial dye concentrations, beginning from
the lowest (C0=20 mg/L) and ending to the highest (C0=1000 mg/L). It is obvious that, the adsorption
Table 3 presents the adsorption capacities of various adsorbents for Reactive Black 5 [37-44]. The
composite material exhibits a significant high capacity, higher than its components, which indicates an
Furthermore, Table 4 presents the thermodynamic parameters resulted from the experimental data
of isotherms at three different temperatures, calculating the respective parameters ( G0, H0, S0). The Gibbs
free energy change ( G0, kJ/mol) of the adsorption process is expressed by the Van’t Hoff equation (Eq.
(7)) [45], with the previous calculation of the equilibrium constant (Kc) (Eq. (8)):
ΔG 0 = − R T ln ( K c ) (7)
Cs
Kc = (8)
Ce
where Cs (mg/L) is the amount adsorbed on solid at equilibrium and R (=8.314 J/mol K) is the universal gas
constant.
The change in entropy ( S0, kJ/mol K) and the heat of adsorption ( H0, kJ/mol) at a constant temperature T
⎛ ΔH0 ⎞ 1 ΔS0
ln(Kc ) = ⎜ − ⎟ + (10)
⎜ R ⎟T R
⎝ ⎠
The values of H0 and S0 were calculated from the slop and intercept of the plot between ln(Kc) versus 1/T.
16
The negative values of ΔG0 showed that the adsorption of RB5 onto all adsorbents used was
0
spontaneous. The positive values of of adsorption indicated an endothermic nature. The latter can be
further explained as a result of the desorption of water molecules, which were previously adsorbed onto the
dye molecule and the adsorption of dye molecules onto the surface of adsorbent material. The positive
values of S0 demonstrated the increased randomness at the solid/liquid interface. The water coordinated
molecules are displaced by dye molecules and consequently gain more translational entropy than is lost by
dye molecules (this occurs during adsorption); the latter results in increased randomness in the dye–
In dyeing technology many additives used, such as salts and surfactants can either accelerate or
retard dye adsorption processes. Sodium chloride, which is often used as a stimulator in dyeing processes
can act in dual mode: (i) it may screen the electrostatic interaction of opposite charges in adsorbents and the
dye molecules, and an increase in salt concentration could decrease the amount of dye adsorbed, (ii) it may
enhance the degree of dissociation of the dye molecules and facilitate the amount of pollutant adsorbed
[35]. So, the ionic strength is another important factor in the dye adsorption process. Fig. 10 presents the
effect of ionic strength on dye adsorption. The behavior of materials containing cross–linked chitosan (Ch
and GO–Ch) showed the same behavior. Increasing the concentration of NaCl from 0.001 to 1 mol/L in dye
solution, a strong augmentation of adsorption capacity is observed reaching ~70 mg/g for Ch and 80 mg/g
for the GO–Ch composite. On the contrary, a deep decrease was observed for GO increasing the salinity
(reduction ~70 mg/g). In the case of GO the excess of ions acted competitively with the charged parts of
dye for the adsorption onto surface of material reducing the electrostatic interactions/forces, while for
chitosan–based materials (Ch, GO–Ch), the adsorption is mainly based on coulombic/electrostatic bonds
which overcome the ions from salts and be sub–helped from the dissociation of dye molecules.
17
3.6. Desorption
The opposite phenomenon of adsorption is called desorption and has a great impact on
decolorization technology. Given the recent industrial trend for regeneration and consequently reuse of
adsorbents, desorption plays an important role for the selection (as adsorbent) of a material. So, a material,
which can easily desorb, is valuable. In this way, the experiments for the pH effect on desorption revealed
that the optimum desorption pH value was alkaline. At pH=12 the desorption of RB5 was 95, 95, and 80%
for GO–Ch, Ch, and GO, respectively. Decreasing the pH of the eluant, desorption reduces. In many dye–
adsorbent systems, desorption phenomenon is just the reverse of that of adsorption. For example, if the
optimum adsorption conditions are alkaline, the optimum desorption ones are acidic, and vice versa.
However, this is does not ever occur, since dye adsorption is more complex phenomenon, based on a
combination of interactions, and not only the classic electrostatic forces (e.g. hydrogen bonding, π–π
4. Conclusions
A graphite oxide/chitosan composite (GO–Ch) was prepared in order to remove Reactive Black 5
dye from aqueous solutions. The GO presented the sheet–like structure, while after the combination with
Ch a rougher surface was presented in GO–Ch, revealing that Ch had been assembled on the surface of GO
layers with a high density. The amorphous–like structure of Ch was not changed by the addition of GO in
the composite material as observed from XRD patterns. According to the same technique, it is concluded
that a significant part of the GO did not exfoliated/reduced on reacting with Ch resulting in an increased
heterogeneity of the GO–Ch composite. A deeper characterization of the prepared composite was also
achieved with other techniques as TGA, DTG, DTA. The interactions between dye and materials are
determined with FTIR spectroscopy, leading to conclusions about the adsorption mechanism of dye onto
composite. The amino groups of chitosan can react with oxygen–containing functional groups of GO in the
following ways: (i) after hydrogen–bonding interactions between the amine molecules and the oxygen–
18
containing functional groups of GO (G–OH…H2N–R); (ii) after protonation of the amine by the weakly
acidic sites of the GO layers (–COO–+H3N–R) with amides and amine carboxylate salts as reactions
products, and (iii) after nucleophilic substitution reactions on the epoxy groups of GO. In the field of
adsorption evaluation of the material prepared, the optimum pH found after adsorption experiments was
pH=2. The equilibrium data were fitted to the Langmuir, Freundlich and L–F model, while the calculated
Qmax for the composite material at 25 oC was 277 mg/g (pH=2). The positive values of H0 suggest the
endothermic nature of the process, while the negative values of G0 suggest that the process is spontaneous
(positive values of S0 confirmed the increased randomness at the solid/liquid interface). Kinetic data were
fitted to the pseudo–first, –second order equations and Brouers–Sotolongo model and the optimum pH for
The Supplementary Information section includes: (i) Fig. SI2, which illustrates the cross–linking
mechanism of chitosan; (ii) Fig. SI1, which demonstrates the differences in FTIR spectra of pure (non
19
References
[1] W.S. Wan Ngah, I.M. Isa, Comparison study of copper ion adsorption on chitosan, dowex A-1, and
[2] F.C. Wu, R.L. Tseng, R.S. Juang, Kinetic modeling of liquid-phase adsorption of reactive dyes and
[3] G. Crini, P.M. Badot, Application of chitosan, a natural aminopolysaccharide, for dye removal from
aqueous solutions by adsorption processes using batch studies: A review of recent literature, Prog. Polym.
[4] V.M. Boddu, K. Abburi, J.L. Talbott, E.D. Smith, R. Haasch, Removal of arsenic (III) and arsenic (V)
from aqueous medium using chitosan-coated biosorbent, Water Res. 42 (2008) 633-642.
[5] A. Bhatnagar, M. Sillanpää, Applications of chitin- and chitosan-derivatives for the detoxification of
water and wastewater - A short review, Adv. Colloid Interface Sci. 152 (2009) 26-38.
[6] S. Stankovich, D.A. Dikin, G.H.B. Dommett, K.M. Kohlhaas, E.J. Zimney, E.A. Stach, R.D. Piner, S.T.
Nguyen, R.S. Ruoff, Graphene-based composite materials, Nature 442 (2006) 282-286.
[7] G.K. Ramesha, A. Vijaya Kumara, H.B. Muralidhara, S. Sampath, Graphene and graphene oxide as
effective adsorbents toward anionic and cationic dyes, J. Colloid Interface Sci. 361 (2011) 270-277.
[8] L. Fan, C. Luo, X. Li, F. Lu, H. Qiu, M. Sun, Fabrication of novel magnetic chitosan grafted with
graphene oxide to enhance adsorption properties for methyl blue, J. Hazard. Mater. 215-216 (2012) 272-
279.
[9] L. Fan, C. Luo, M. Sun, X. Li, F. Lu, H. Qiu, Preparation of novel magnetic chitosan/graphene oxide
composite as effective adsorbents toward methylene blue, Bioresour. Technol. 114 (2012) 703-706.
[10] Y.L. Wang, P. Gao, L.H. Huang, X.J. Wu, Y.L. Liu, Graphite oxide: Preparation and removal ability
[11] H. Zollinger, in: (Eds.), Color Chemistry: Synthesis, Properties and Applications of Organic Dyes and
Pigments, 1987,
20
[12] M. Rinaudo, Chitin and chitosan: Properties and applications, Prog. Polym. Sci. 31 (2006) 603-632.
[13] N.K. Lazaridis, G.Z. Kyzas, A.A. Vassiliou, D.N. Bikiaris, Chitosan derivatives as biosorbents for
[14] W.S. Hummers Jr, R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc. 80 (1958) 1339.
[15] S. Lagergren, About the theory of so-called adsorption of soluble substances, Handlingar 24 (1898) 1-
39.
[16] Y.S. Ho, J.C.Y. Ng, G. McKay, Kinetics of pollutant sorption by biosorbents: Review, Sep. Purif.
[17] S.H. Chien, W.R. Clayton, Application of Elovich equation to the kinetics of phosphate release and
[18] D.L. Sparks, Kinetics of Reaction in Pure and Mixed Systems, in: D.L. Sparks (Eds.), Soil Physical
Chemistry, CRC Press, Boca Raton, Florida, U.S.A., 1986, pp. 63-145.
[19] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and platinum, J. Am. Chem.
[20] H. Freundlich, Over the adsorption in solution, Z. Phys. Chem. 57 (1906) 385-470.
[21] C. Tien, Adsorption Calculations and Modeling, Butterworth-Heinemann, Boston, U.S.A., 1994.
[22] L. Liu, C. Li, C. Bao, Q. Jia, P. Xiao, X. Liu, Q. Zhang, Preparation and characterization of
chitosan/graphene oxide composites for the adsorption of Au(III) and Pd(II), Talanta 93 (2012) 350-357.
[23] M. Seredych, R. Pietrzak, T.J. Bandosz, Role of graphite oxide (GO) and polyaniline (PANI) in NO 2
[24] M. Seredych, A.V. Tamashausky, T.J. Bandosz, Graphite oxides obtained from porous graphite: The
role of surface chemistry and texture in ammonia retention at ambient conditions, Adv. Funct. Mater. 20
(2010) 1670-1679.
[25] H.K. Jeong, H.J. Noh, J.Y. Kim, M.H. Jin, C.Y. Park, Y.H. Lee, X-ray absorption spectroscopy of
21
[26] H.H. Seung, Thermal Reduction of Graphene Oxide, in: S. Mikhailov (Eds.), Physics and Applications
[27] A.B. Bourlinos, D. Gournis, D. Petridis, T. Szabó, A. Szeri, I. Dékány, Graphite oxide: Chemical
reduction to graphite and surface modification with primary aliphatic amines and amino acids, Langmuir 19
(2003) 6050-6055.
[28] Z.H. Liu, Z.M. Wang, X. Yang, K. Ooi, Intercalation of organic ammonium ions into layered graphite
[29] Y. Guo, X. Sun, Y. Liu, W. Wang, H. Qiu, J. Gao, One pot preparation of reduced graphene oxide
(RGO) or Au (Ag) nanoparticle-RGO hybrids using chitosan as a reducing and stabilizing agent and their
derivative with a complexing agent and its adsorption studies toward metal(II) ions, Carbohydr. Res. 344
(2009) 1632-1638.
[33] K.S.V. Krishna Rao, I. Chung, C.-S. Ha, Synthesis and characterization of poly(acrylamidoglycolic
acid) grafted onto chitosan and its polyelectrolyte complexes with hydroxyapatite, React. Funct.l Polym. 68
(2008) 943-953.
[34] G.Z. Kyzas, N.K. Lazaridis, Reactive and basic dyes removal by sorption onto chitosan derivatives, J.
[35] G. Crini, Non-conventional low-cost adsorbents for dye removal: A review, Bioresour. Technol. 97
(2006) 1061-1085.
[36] K.R. Hall, L.C. Eagleton, A. Acrivos, T. Vermeulen, Pore- and solid-diffusion kinetics in fixed-bed
22
adsorption under constant-pattern conditions, Ind. Eng. Chem. Fundam. 5 (1966) 212-223.
[37] G. Gibbs, J.M. Tobin, E. Guibal, Influence of Chitosan Preprotonation on Reactive Black 5 Sorption
[38] K.Z. Elwakeel, M. Rekaby, Efficient removal of Reactive Black 5 from aqueous media using glycidyl
methacrylate resin modified with tetraethelenepentamine, J. Hazard. Mater. 188 (2011) 10-18.
[39] A.W.M. Ip, J.P. Barford, G. McKay, A comparative study on the kinetics and mechanisms of removal
of Reactive Black 5 by adsorption onto activated carbons and bone char, Chem. Eng. J. 157 (2010) 434-
442.
[40] O. Ozdemir, B. Armagan, M. Turan, M.S. Çelik, Comparison of the adsorption characteristics of azo-
[41] R. Patel, S. Suresh, Kinetic and equilibrium studies on the biosorption of reactive black 5 dye by
[42] B. Arma an, O. Özdemir, M. Turan, M.S. Çelik, The removal of reactive azo dyes by natural and
[43] Z. Eren, F.N. Acar, Adsorption of Reactive Black 5 from an aqueous solution: equilibrium and kinetic
[44] J.F. Osma, V. Saravia, J.L. Toca-Herrera, S.R. Couto, Sunflower seed shells: A novel and effective
low-cost adsorbent for the removal of the diazo dye Reactive Black 5 from aqueous solutions, J. Hazard.
[45] J.M. Smith, H.C. Van Ness, in: (Eds.), Introduction to Chemical Engineering Thermodynamics,
[46] M.R. Unnithan, T.S. Anirudhan, The kinetics and thermodynamics of sorption of chromium(VI) onto
the iron(III) complex of a carboxylated polyacrylamide-grafted sawdust, Ind. Eng. Chem. Res. 40 (2001)
2693-2701.
23
24
Figure List
Fig. 2. X–ray diffraction (XRD) patterns of Ch, GO, and GO–Ch (Inset: XRD patterns of
graphite and GO).
Fig. 6. FT–IR spectra of adsorbents before and after adsorption of RB5: (A) GO; (B) dye–
loaded GO; (C) Ch; (D) dye–loaded Ch; (E) GO–Ch; (F) dye loaded GO–Ch.
Fig. 7. Effect of initial pH on adsorption of RB5 onto Ch, GO, and GO–Ch (pH=2–12; 250
mg/L dye concentration; 1 g/L adsorbent; T=25 °C; 160 rpm; 0.001 mol/L NaCl; 24 h
contact).
Fig. 8. Effect of contact time on adsorption RB5 onto Ch, GO, and GO–Ch (pH=2; 250 mg/L
initial dye concentration; 1 g/L adsorbent; T=25 °C; 160 rpm; 0.001 mol/L NaCl; 0–24
h contact), fitted to (a) pseudo–first order, (b) pseudo–second order, and (c) Elovich
equation.
Fig. 9. Effect of initial dye concentration on adsorption of RB5 onto Ch, GO, and GO–Ch
(pH=2; 0–1000 mg/L dye concentration; 1 g/L adsorbent; T=25, 45, 65 °C; 160 rpm;
0.001 mol/L NaCl; 24 h contact).
Fig. 10. Effect of ionic strength on adsorption RB5 onto Ch, GO, and GO–Ch (pH=2; 250 mg/L
initial dye concentration; 1 g/L adsorbent; T=25 °C; 160 rpm; 0.001, 0.1 and 1 mol/L
NaCl; 24 h contact).
25
FIGURE CAPTIONS
NH2 OH O H2 H
H2 H2 O 2
NaO3SO C C S N N N N S C C OSO3Na
O O
HO3S SO3H
26
1400
Graphite
1200
Intensity (counds/sec)
1000
800 GO
600 GO-Ch 10 20 30 40 50
2θ (deg)
400
200
Ch GO
5 10 15 20 25 30 35 40 45 50
2θ (deg)
Fig. 2. X–ray diffraction (XRD) patterns of Ch, GO, and GO–Ch (Inset: XRD patterns of graphite and
GO).
27
(a) (b)
28
0.9 100
90
0.8
Weight loss derivative (%/ C)
80
7%
o
Weight (%)
70
0.7 26% 28% 40%
60 Ch
50
0.6 40
30 GO-Ch
0.5 GO-Ch 20
GO
0 200 400 600 800
0.4 GO o
Temperature ( C)
0.3
0.2 Ch
0.1
0.0
0 50 100 150 200 250 300 350 400 450 500
o
Temperature ( C)
Fig. 4. Differential thermogravimetry (DTG) curves in nitrogen (Inset: Thermogravimetric analysis (TGA)
29
4
3
Temperature Difference ( C)
o
2
1 GO-Ch
0
Ch
-1
-2
GO-Ch
-3
-4
0 100 200 300 400 500 600 700 800 900
o
Temperature ( C)
30
A
Transmitance (%) B
Fig. 6. FT–IR spectra of adsorbents before and after adsorption of RB5: (A) GO; (B) dye–loaded GO; (C)
Ch; (D) dye–loaded Ch; (E) GO–Ch; (F) dye loaded GO–Ch.
31
100 14
Ch GO GO-Ch
90
12
80
70 10
Adsorption (%)
60
8
pH,fin
50
6
40
30 4
20
2
10
0 0
0 2 4 6 8 10 12 14
pH,in
Fig. 7. Effect of initial pH on adsorption of RB5 onto Ch, GO, and GO–Ch (pH=2–12; 250 mg/L dye
concentration; 1 g/L adsorbent; T=25 °C; 160 rpm; 0.001 mol/L NaCl; 24 h contact).
32
250
225
200
175
150
Qt (mg/g)
125
100
75
50
Ch
25 GO
GO-Ch
0
0 4 8 12 16 20 24 28
t (h)
Fig. 8. Effect of contact time on adsorption RB5 onto Ch, GO, and GO–Ch (pH=2; 250 mg/L initial dye
concentration; 1 g/L adsorbent; T=25 °C; 160 rpm; 0.001 mol/L NaCl; 0–24 h contact).
33
350
Ch
300 GO
GO-Ch
Qe (mg/g) 250
200
150
100
50
0
0 50 100 150 200 250 300
Ce (mg/L)
Fig. 9. Effect of initial dye concentration on adsorption of RB5 onto Ch, GO, and GO–Ch (pH=2; 0–1000
mg/L dye concentration; 1 g/L adsorbent; T=25 °C; 160 rpm; 0.001 mol/L NaCl; 24 h contact).
34
320 Ch GO GO-Ch
300
280
Qe (mg/g) 260
240
220
200
180
160
140
120
100
0.001 0.01 0.1 1
NaCl (mol/L)
Fig. 10. Effect of ionic strength on adsorption capacity of Ch, GO, and GO–Ch (pH=2; 250 mg/L initial
dye concentration; 1 g/L adsorbent; T=25 °C; 160 rpm; 0.001, 0.01, 0.1 and 1 mol/L NaCl; 24 h contact).
35
Table 1
Kinetic constants for the adsorption of RB5 onto Ch, GO, and GO–Ch.
⎡ ⎛ ⎞⎤ 1/ ( n −1)
Q t = Q e (1 − e − k1t ) Q t ( n,a ) = Qe ⎡1 − (1 + Qe n −1 ( n − 1) k n,a t a )
1 ⎤
Q t = Qe ⎢1 − ⎜ ⎟⎥
⎣ ⎝ 1 + Qe k 2 t ⎠ ⎦ ⎣⎢ ⎦⎥
k1 R2 k2 R2 n a kn,a R2
Adsorbent min–1 min–1(mg/g)–1 min–a(mg/g)1–n
Ch 0.987 0.983 0.010 0.943 1.462 1.253 0.213 0.999
GO 2.067 0.512 0.049 0.421 1.197 0.959 0.517 0.966
GO–Ch 0.765 0.979 0.006 0.988 1.899 1.188 0.011 0.988
k,, k2, kn,a are the rate constants for the pseudo–first, pseudo –second order and Brouers–Sotolongo kinetic models, respectively; n is the fractional
reaction order of Brouers–Sotolongo model; a “fractal time” parameter of Brouers–Sotolongo model.
36
Table 2
Equilibrium parameters for the adsorption of RB5 onto Ch, GO and GO–Ch at 25, 45, and 65 oC.
Langmuir equation Freundlich equation Langmuir–Freundlich (L–F) equation
Qmax ( K LFCe )
b
Q max K L Ce 1/n
Qe = Q e = K FC Qe =
1+ ( K LFCe )
e b
1+K L Ce
37
Table 3
Comparative table presenting the maximum adsorption capacities of different
materials after adsorption of RB5.
Adsorbent Qmax (mg/g) Reference
Protonated chitosan 550 [37]
Glycidyl methacrylate resin 484 [38]
BACX6 473 [39]
BACX2 286 [39]
Graphite oxide/chitosan composite 277 This study
Chitosan 224 This study
Graphite Oxide 205 This study
F400 198 [39]
Bone char 160 [39]
Sepiolite 121 [40]
Aspergillus foetidus biomass 92 [41]
Modified zeolites 61 [42]
Activated Carbon 59 [43]
Fly Ash 8 [43]
Sunflower seed shells 1 [44]
38
Table 4
Thermodynamic parameters for the adsorption of RB5 onto GO, Ch, and GO–Ch.
C0 T Qe Kc G0 H0 S0
Adsorbent (mg/L) (K) (mg/g) (kJ/mol) (kJ/mol) (kJ/mol K)
Ch 20 298 18.97 18.49 –7.23
318 19.00 19.00 –7.78 +5.35 +0.042
338 19.20 24.00 –8.93
80 298 76.92 24.99 –7.97
318 77.25 28.09 –8.82 +8.67 +0.056
338 77.95 37.99 –10.22
300 298 240.49 4.04 –3.46
318 250.75 5.09 –4.30 +10.53 +0.047
338 261.01 6.69 –5.34
GO 20 298 19.67 59.61 –10.13
318 19.72 70.43 –11.25 +10.53 +0.069
338 19.80 99.00 –12.91
80 298 60.61 3.12 –2.82
318 65.14 4.39 –3.91 +53.40 +0.186
338 78.13 41.81 –10.49
300 298 139.07 0.86 0.36
318 175.04 1.40 –0.89 +21.98 +0.072
338 213.73 2.48 –2.55
GO–Ch 20 298 19.72 65.67 –10.37
318 19.81 99.00 –12.15 +37.29 +0.159
338 19.95 399.00 –16.83
80 298 75.57 17.06 –7.03
318 77.05 26.12 –8.63 +22.76 +0.100
338 78.46 50.95 –11.05
300 298 227.11 3.11 –2.81
318 261.15 6.50 –4.95 +27.41 +0.102
338 275.98 11.50 –6.86
39
Highlights
40