Vous êtes sur la page 1sur 16

Journal of Colloid and Interface Science 512 (2018) 575–590

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


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

Regular Article

Synthesis and application of a new carboxylated cellulose derivative.


Part III: Removal of auramine-O and safranin-T from mono- and
bi-component spiked aqueous solutions
Filipe Simões Teodoro a, Megg Madonyk Cota Elias a, Gabriel Max Dias Ferreira a,
Oscar Fernando Herrera Adarme b, Ranylson Marcello Leal Savedra c, Melissa Fabíola Siqueira c,
Luis Henrique Mendes da Silva d, Laurent Frédéric Gil a, Leandro Vinícius Alves Gurgel a,⇑
a
Grupo de Físico-Química Orgânica, Departamento de Química, Instituto de Ciências Exatas e Biológicas, Universidade Federal de Ouro Preto, Campus Universitário Morro do
Cruzeiro, s/n°, Bauxita, 35400-000 Ouro Preto, Minas Gerais, Brazil
b
Laboratório de Química Tecnológica e Ambiental, Departamento de Química, Instituto de Ciências Exatas e Biológicas, Universidade Federal de Ouro Preto, Campus
Universitário Morro do Cruzeiro, s/n°, Bauxita, 35450-000 Ouro Preto, Minas Gerais, Brazil
c
Laboratório de Polímeros e Propriedades Eletrônicas de Materiais, Departamento de Física, Universidade Federal de Ouro Preto, Campus Universitário Morro do Cruzeiro,
s/n°, Bauxita, 35400-000 Ouro Preto, Minas Gerais, Brazil
d
Grupo de Química Verde Coloidal e Macromolecular, Departamento de Química, Centro de Ciências Exatas e Tecnológicas, Universidade Federal de Viçosa, Av. P. H. Rolfs, s/n°,
36570-000 Viçosa, Minas Gerais, Brazil

g r a p h i c a l a b s t r a c t

Adsorption Desorption Re-adsorption


HO O
OH O

Auramine-O qe-re= 2.67 mmol g-1


+ O
Py:DMA (1:1, v/v) Edes= 44.61 % Ere-ads= 77.90 %
OH
HO
OH
O
100 oC; 1 h; 300 rpm HO

n O Trimellitic O
Cellulose anhydride O
O O pH 4.5 qe = 2.628 mmol g-1
pH 7 qe = 5.172 mmol g-1
RO O
OR

n Binary System [HCl] = 0.01mol L-1


Cellulose modified 130 rpm
with Trimellitic Anhydride (CTA)
3 h, 25°C

AO qe = 1.051 mmol g-1


pH 7
ST qe = 3.258 mmol g-1

qe-re= 3.64mmol g-1


Edes= 78.87 % Ere-ads= 60.71 %
Celulose modified Safranin-T
Cellulose
with Trimellitic Anhydride (CTA)
pH 4.5 qe = 3.443 mmol g-1
pH 7 qe = 4.019 mmol g-1

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

Article history: In the third part of this series of studies, the adsorption of the basic textile dyes auramine-O (AO) and
Received 21 July 2017 safranin-T (ST) on a carboxylated cellulose derivative (CTA) were evaluated in mono- and bi-
Revised 21 October 2017 component spiked aqueous solutions. Adsorption studies were developed as a function of solution pH,
Accepted 23 October 2017
contact time, and initial dye concentration. Adsorption kinetic data were modeled by monocomponent
Available online 25 October 2017
kinetic models of pseudo-first- (PFO), pseudo-second-order (PSO), intraparticle diffusion, and Boyd, while
the competitive kinetic model of Corsel was used to model bicomponent kinetic data. Monocomponent
Keywords:
adsorption equilibrium data were modeled by the Langmuir, Sips, Fowler-Guggenhein, Hill de-Boer,
Adsorption
Auramine-O
and Konda models, while the IAST and RAST models were used to model bicomponent equilibrium data.
Safranin-T Monocomponent maximum adsorption capacities for AO and ST at pH 4.5 were 2.841 and 3.691 mmol
Desorption g1, and at pH 7.0 were 5.443 and 4.074 mmol g1, respectively. Bicomponent maximum adsorption
Isothermal titration calorimetry capacities for AO and ST at pH 7.0 were 1.230 and 3.728 mmol g1. Adsorption enthalpy changes

⇑ Corresponding author.
E-mail addresses: legurgel@iceb.ufop.br, legurgel@yahoo.com.br (L.V.A. Gurgel).

https://doi.org/10.1016/j.jcis.2017.10.083
0021-9797/Ó 2017 Elsevier Inc. All rights reserved.
576 F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590

Thermodynamics (DadsH) were obtained using isothermal titration calorimetry. The values of DadsH ranged from 18.83 to
Molecular quantum mechanics 5.60 kJ mol1, suggesting that physisorption controlled the adsorption process. Desorption and re-
adsorption of CTA was also evaluated.
Ó 2017 Elsevier Inc. All rights reserved.

1. Introduction ions from mono- and multi-component aqueous solutions, in addi-


tion to CTA’s high capacity for basic dye adsorption compared to
Contamination of ground and surface water by synthetic dyes previously studied adsorbents (Supplementary Table 1), motivated
due to the discharge of untreated or partially treated industrial our research group to begin adsorption studies for the removal of
wastewater is a serious environmental problem and is considered dyes from aqueous solutions. As pointed out by Gusmão et al.
to be a threat to humans and aquatic organisms [1,2]. Synthetic [12], the use of the same adsorbent for various environmental
dyes have various industrial uses including in cosmetics, food pro- applications is desirable because it makes its production econom-
cessing, leather tanning, rubber, paper, plastics, printing and dye ically feasible.
manufacturing, and textiles [1,3]. In the third part of this series of studies, the removal of the basic
The discharge of synthetic dyes into the hydrosphere is prob- dyes auramine-O (AO) and safranin-T (ST) using the CTA adsorbent
lematic primarily due to dyes’ potential to color water and reduce was assessed as a function of solution pH, contact time (kinetics),
sunlight penetration into water, as well as dyes’ resistance to pho- and initial dye concentration in spiked mono- and bi-component
tochemical and biological degradation [1,4]. In addition, synthetic aqueous solutions. Dye removal was also studied at two pH values
dyes and their degradation products include mutagenic com- (i.e. 4.5 and 7.0). A suitable desorption process was developed to
pounds, carcinogenic compounds, and compounds that may evaluate the regeneration and reusability of the CTA adsorbent.
increase the chemical oxygen demand (COD) and biochemical oxy- The enthalpy of the adsorption process was also evaluated and
gen demand (BOD) of water resources [1,3]. The complex aromatic assessed by isothermal titration calorimetry, and the removal
chemical structures of synthetic dyes make them very stable to mechanism of both dyes was evaluated based on thermodynamic
heat, oxidizing agents, photochemical degradation, and biodegra- data and molecular quantum mechanics calculations.
dation, which makes their removal from industrial wastewaters
difficult [1,3]. 2. Material and methods
Auramine-O and safranin-T are classified as basic dyes [5],
which can be used to dye wool, silk, acrylic fibers, and leather in 2.1. Material
textile industries [6]. Auramine-O can cause acute oral toxicity,
carcinogenicity, cytotoxicity, DNA damage, genotoxicity, and Auramine-O (MW = 303.83 g mol1; C17H21N3HCl; C.I. 41000;
mutagenicity [7]. Safranin-T can cause carcinogenicity, mitochon- kmax = 434 nm; purity  100%) and safranin-T (MW = 350.84 g
drial toxicity, mutagenicity, and nucleic acid damage [7]. According mol1; C20H19N4Cl; C.I. 50240; kmax = 530 nm; purity  100%) were
to Terinte et al. [8], most conventional dyeing takes place in devel- purchased from Vetec (Brazil). Anhydrous citric acid (C6H8O7; 99%)
oping countries, where environmental legislation is often insuffi- and trisodium citrate (Na3C6H5O72H2O; 99%) were purchased
cient and/or not implemented properly. A survey describing the from Synth (Brazil). Dye solutions used in the adsorption studies
percentages of unfixed dyes that may be discharged in the effluent were prepared in buffer solutions consisting of 0.05 mol L1 citric
for different dyes, fabrics, and applications was recently reported acid/sodium monocitrate (pH from 2.13 to 3.13), 0.05 mol L1
by Terinte et al. [8]. sodium monocitrate/sodium dicitrate (pH from 4.13 to 5.15), and
Three categories of treatment technologies have been applied to 0.05 mol L1 sodium dicitrate/sodium tricitrate (pH from 6.13 to
effluent containing synthetic dyes: physical, chemical, and biolog- 7.13). Other materials and chemicals used in this investigation
ical [3,9]. Physical treatments include adsorption by activated car- are described in the first part of this series of studies [10].
bon, membrane filtration and separation, ion-exchange,
irradiation, and electrokinetic coagulation. Chemical treatments 2.2. Adsorption experiments
include oxidative processes, photochemical destruction, and elec-
trochemical destruction. Biological treatments include aerobic Monocomponent adsorption studies of AO and ST on the CTA
and anaerobic treatment, among others [1,3,9]. Despite the devel- adsorbent were performed at two pH values (i.e. 4.5 and 7.0,) while
opment and application of technologies for treatment of wastewa- bicomponent adsorption studies of AO and ST on the CTA adsor-
ters containing synthetic dyes, cost effective, efficient, and rapid bent were only carried out at pH 7.0, which was the pH of highest
water treatment at the commercial level is still a challenging prob- adsorption capacity for both dyes. All adsorption studies were per-
lem for the scientific community [1]. Although adsorption can treat formed in duplicate.
large flow rates, producing high quality effluents [1], if the adsor-
bent material cannot be cleaned by desorption and reused, the
2.2.1. Adsorption of dyes onto CTA as a function of solution pH
applicability and usefulness of this treatment method is limited.
For a typical monocomponent experiment, CTA samples (20.0
Thus, research to develop suitable adsorbents with high adsorption
mg) weighed in cylindrical glasses (1.8 mm height  2.2 mm
capacity, good economic viability, versatility for removal of various
diameter) were added to 250 mL Erlenmeyer flasks containing
organic and inorganic pollutants, and better regeneration capacity
100.0 mL of buffered dye solution of a known concentration
is still needed.
(0.932 mmol L1) at pH values varying from 2.13 to 7.14. Flasks
In the first and second parts of this series of studies [10,11], the
were stirred at 130 rpm and 25 °C in a shaker incubator (Marconi,
synthesis and application of a new adsorbent material (CTA) using
model MA-830) for 24 h. Then, the suspensions were centrifuged
cellulose as a solid support was described in detail, as well as the Ò
(Excelsa II centrifuge, model 206 BL) at 3600 rpm for 10 min to
adsorption of Co2+, Cu2+, and Ni2+ from mono- and multi-
separate the solid and liquid phases. The absorbance of the super-
component aqueous solutions. The high adsorption capacity and
natant phase was measure on a UV–Vis spectrophotometer
selectivity exhibited by the CTA adsorbent for removal of metal
(Biospectro, model SP-220) at 434 nm for AO and 530 nm for ST.
F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590 577

Calibration curves were constructed to determine the dye concen- 1 ðC i  C e ÞV


qe =ðmmol g Þ ¼ ð5Þ
tration in the supernatant phase. wCTA
For bicomponent experiments, the procedures described above
where qe (mmol g1) is the adsorption capacity of dye on the CTA
were used, except that buffered dye solutions contained both dyes
adsorbent at equilibrium and Ci and Ce (mmol L1) are the dye solu-
at known concentrations (0.93 mmol L1). In a bicomponent dye
tion concentrations initially and at equilibrium, respectively.
solution, the concentrations of A (AO) and B (ST), measured at k1
For bicomponent experiments, 100.0 mL of buffered dye solu-
and k2, respectively, to give absorbances of A1 and A2, were deter-
tions at pH 7.0 with dye concentrations varying from 0.17 to
mined using Eqs. (1) and (2) [13]:
1.06 mmol L1 for AO and 0.19 to 1.16 mmol L1 for ST were added
kB2 A1  kB1 A2 to the flasks, which were stirred at 130 rpm and 25 °C for 24 h.
CA ¼ ð1Þ
kA1 kB2  kA2 kB1 Other procedures were the same as described in Section 2.2.1.

kA1 A2  kA2 A1 2.3. Desorption experiments


CB ¼ ð2Þ
kA1 kB2  kA2 kB1
CTA adsorbent was loaded with AO or ST using the same exper-
where kA1 (0.12957), kB1 (0.01471), kA2 (0.0), and kB2 (0.10750) are
imental procedures described in Section 2.2.1. Dye solutions of
the calibration constants for components A and B at the wave-
0.87 mmol L1 for AO and 0.99 mmol L1 for ST at pH 7.0 were
lengths k1 and k2, respectively.
used. After loading the CTA adsorbent with a dye, the CTA adsor-
bent was separated by a single filtration step and was rinsed with
2.2.2. Adsorption of dyes onto CTA as a function of contact time
an excess of distilled water to remove unadsorbed dye molecules.
In a typical run for monocomponent experiments, Erlenmeyer
Then, the CTA adsorbent loaded with dye was dried in an oven at
flasks of 250 mL containing 100.0 mL of buffered dye solution of
90 °C for 3 h.
a known concentration (0.932 mmol L1) at pH values of 4.5 and
7.0 were thermostated in a shaker incubator at 25 °C for 1 h. Then,
2.3.1. Desorption of the CTA adsorbent as a function of desorption
CTA samples of 20.0 mg weighed in cylindrical glasses were added
agent type
to each flask and stirred at 130 rpm and 25 °C for time intervals
Two desorption agents (HCl and HNO3) at two concentrations
from 5 to 1920 min. Adsorption capacity was calculated using Eq.
(0.01 and 0.1 mol L1) were evaluated for desorption of CTA adsor-
(3).
bent loaded with AO or ST. Samples of CTA loaded with a dye (20.0
1 ðC 0  C e ÞV mg) weighed into cylindrical glasses were added to 100 mL Erlen-
qt =ðmmol g Þ ¼ ð3Þ
wCTA meyer flasks containing 20.0 mL of desorption solution and stirred
at 130 rpm and 25 °C for 16 h. Solid and liquid phases were sepa-
where qt (mmol g1) is the adsorption capacity of AO or ST on the
rated by centrifugation. The pH of the supernatant was adjusted
CTA adsorbent at time t, V (L) is the volume of dye solution, C0
to the pH at which the calibration curve was obtained by dropwise
and Ct (mmol L1) are the dye solution concentrations initially
addition of aqueous 0.1 mol L1 NaOH. Absorbance was measured
and at time t, respectively, and wCTA (g) is the weight of the CTA
on a UV–Vis spectrophotometer as described in Section 2.2.1. All
adsorbent.
desorption experiments were performed in triplicate. Desorption
For bicomponent experiments, the same experimental proce-
efficiency was calculated using Eq. (6) [10].
dures described above were used, except that buffered dye solu-  
tions contained both dyes at known concentrations (0.932 mmol C e;des V
Edes =% ¼  100 ð6Þ
L1) at pH value 7.0 were added to the Erlenmeyer flasks. The flaks Q T;max w0CTA
were stirred for different time intervals from 5 to 1800 min. Other
where Edes (%) is the desorption efficiency, Ce,des (mg L1) is the
procedures were the same as described in Section 2.2.1.
equilibrium dye concentration in the desorption solution, QT,max
(mg g1) is the maximum adsorption capacity determined by load-
2.2.3. Adsorption of dyes onto CTA as a function of adsorbent dosage
ing a dye (AO or ST) on the CTA adsorbent, and w0 CTA (g) is the
CTA samples of 10.0, 20.0, 30.0, 40.0, 50.0, 60.0 and 70.0 mg
weight of the CTA adsorbent in wCTA,dye, calculated using Eq. (7)
weighed in cylindrical glasses were added to 250 mL Erlenmeyer
[10].
flasks containing 100.0 mL of buffered dye solution of a known
concentration (0.932 mmol L1) at pH 7.0. The flaks were stirred wCTA;dye
w0CTA =g ¼ Q  ð7Þ
at 130 rpm and 25 °C for 24 h. Other procedures were the same T;max
þ1
1000
as described in Section 2.2.1. The dye removal percentage was cal-
culated using Eq. (4).
  2.3.2. Desorption of the CTA adsorbent as a function of time
C0  Ce
R=% ¼  100 ð4Þ Dried samples of CTA loaded with a dye (20.0 mg) weighed into
C0
cylindrical glasses were added to 250 mL Erlenmeyer flasks con-
where C0 and Ce (mmol L1) are the dye solution concentrations ini- taining 100.0 mL of aqueous 0.01 mol L1 HCl solution and stirred
tially and at equilibrium, respectively. at 130 rpm and 25 °C for different time intervals (1, 3, 6, 12, and 24
h). Other procedures were the same described in Section 2.3.1.
2.2.4. Adsorption of dyes onto CTA as a function of initial dye
concentration 2.4. Re-adsorption experiments
In typical monocomponent experiments, CTA samples of 20.0
mg weighed in cylindrical glasses were added to 250 mL Erlen- Dried samples of CTA adsorbent (20.0 mg) from the desorption
meyer flasks containing 100.0 mL of buffered dye solutions at pH study weighed into cylindrical glasses were added to 250 mL
values of 4.5 and 7.0 with dye concentrations varying from 0.21 Erlenmeyer flasks containing 100.0 mL of a dye solution (0.932
to 2.2 mmol L1 for AO and 0.25 to 1.38 mmol L1 for ST. The flasks mmol L1). Other procedures were the same described in Sec-
were stirred at 130 rpm and 25 °C for 24 h. The adsorption capacity tion 2.2.1. All re-adsorption experiments were performed in tripli-
was calculated using Eq. (5). cate. The re-adsorption efficiency was calculated using Eq. (8) [10].
578 F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590

 
Q e;reads where ce is the activity coefficient at equilibrium and b (L mol1) is
Ereads =% ¼  100 ð8Þ
Q T;max the Langmuir constant. The activity coefficients of the present study
were calculated using the approximations of Guntelberg [15] and
where Qe,re-ads (mg g1) is the adsorption capacity of the CTA adsor- Davies [16] (Supplementary Material), as suggested by Liu [14]
bent after the desorption process and QT,max (mg g1) is the adsorp- and Ghosal and Gupta [17]. For AO and ST isotherms at pH 4.5,
tion capacity of the CTA adsorbent before the desorption process. the values of the Langmuir constant used in the calculations were
Qe,re-ads is calculated using Eq. (9) [10]. those of the first plateaus (b1).
w0dye þ w00dye
Q e;reads =ðmg g1 Þ ¼ ð9Þ 2.7. Molecular quantum mechanics calculations
wCTA
0
where w dye (mg) is the weight of dye not desorbed from the CTA To provide insight into the structural and electronic behavior of
adsorbent after desorption and w’0 dye (mg) is the weight of dye auramine-O and safranin-T, as well as the monomer model of the
adsorbed on the CTA adsorbent after re-adsorption. carboxylated cellulose derivative (CTA) (Fig. 1), density functional
theory (DFT) calculations were carried out, implemented in the
2.5. Isothermal titration calorimetry electronic structure program package ORCA [18]. The effects of
increased acidity on both dyes were evaluated from calculations
Calorimetry experiments to determine the changes in enthalpy of dyes’ different ionization states. Becke’s three-parameter hybrid
of dye adsorption on CTA were performed in a TAM III isothermal function was used with the non-local correlation of the Lee-Yang-
titration nanocalorimeter (TA instruments, EUA) controlled by Parr (B3LYP) method in the gas phase [19–22], as this method has
TAM assistantTM dedicated software. A concentrated dye solution been used successfully for a variety of organic molecules [23–26].
(330 mg mL1 for both ST and AO) prepared in citrate buffer was This functional was used with the Pople’s basis set 6-31G(2d,2p),
titrated into the reaction cell that initially contained (i) 2.7 mL of which has polarization functions to make it more flexible [27].
citrate buffer or (ii) 2.7 mL of citrate buffer plus 1.35 mg of CTA. Frontier molecular orbitals (FMOs) and molecular electron density
A volume of 2.7 mL of buffer filled the reference cell. Injections (MED) maps were also calculated at the same level of theory. All
of 40 lL of the dye solution were carried out stepwise using a molecular graphics were built using the UCSF Chimera visualiza-
500 lL Hamilton syringe controlled by a piston pump. The time tion package [28].
interval between two consecutive injections was 40 min. The reac-
tion cell was stirred at 180 rpm with a helix stirrer during experi-
3. Results and discussion
ments. Solutions were degassed for 10 min before experiments.
Measurements were performed at pH 4.5 and 7.0. The heat, in kJ,
3.1. Synthesis and characterization of the CTA adsorbent
resulting from each injection i was determined by integration of
the plots of potency against time provided by the equipment.
A detailed characterization of the CTA adsorbent was described
The adsorption enthalpy change (DadsH), in kJ mol1, was obtained
in the first part of this series of studies [10]. For this study, it is
from Eq. (10).
helpful to summarize the amount of free carboxylic acid groups
X
N (nCOOH: 6.81 ± 0.02 mmol g1), the PZC (2.80 ± 0.09), and the textu-
ðqi;int  qi;dil Þ ral properties of the CTA adsorbent (maximum pore diameter of
i¼1
Dads H ¼ ð10Þ 570.95 ± 38.25 Å) [10].
X
N
ni
i¼1 3.2. Theoretical analysis

where qi,int and qi,dil are the energies absorbed or released in the
3.2.1. Model of the CTA adsorbent
reaction cell with and without CTA, respectively, at the ith injection
The CTA adsorbent’s electronic properties were estimated from
of the dye solution, and ni is the amount of dye, in mol, that was
the oligomer model shown in Fig. 1. Only the monomer was ana-
adsorbed on the adsorbent for the same injection. The adsorption
lyzed because the other bonds along the main chain are saturated,
isotherms were used to obtain values of ni. A control experiment
and therefore have minor effects on the electronic structure
adding only buffer to buffer containing 1.35 mg of CTA was per-
results.
formed, and the signal measured was of the same magnitude as
equipment noise. Experiments were carried out at 25.0000 ±
0.0001 °C in duplicate.

2.6. Calculation of free energy of adsorption

Changes in free energy of adsorption (DadsG°) were determined


in a similar way as described in the first part of this series of stud-
ies [10], with the exception of calculating the thermodynamic
equilibrium constant, Ka [14]. Thus, DadsG° values were calculated
using Eq. (11).
1
Dads G =ðkJ mol Þ ¼ RT ln K a ð11Þ
1
where R is the gas constant (8.314 kJ mol ), T (K) is the absolute
temperature, and Ka (dimensionless) is the thermodynamic equilib-
rium constant. Values of Ka for the adsorption system were calcu-
lated using Eq. (12).
Fig. 1. Fully optimized CTA oligomer model, calculated using the level of theory
b 1 B3LYP/6-31(2d,2p). Oxygen atoms are depicted by red, carbon by light brown, and
Ka ¼ ð1 mol L Þ ð12Þ hydrogen by white. (For interpretation of the references to colour in this figure
ce
legend, the reader is referred to the web version of this article.)
F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590 579

Fig. 2 shows that the inner atoms in both rings form a positive CTA dosage due to the decrease in the adsorbate-to-adsorbent ratio
site (blue-cyan in the CTA MED map), including the ester and car- [29]. Based on these results, subsequent adsorption experiments
boxyl groups. In addition, the highest electron density (red areas) is were carried out with a CTA dosage of 0.2 g L1, because greater
localized outside the rings on the oxygen atoms of ether and car- dosages led to significant decreases in the adsorption capacity of
boxyl groups. The frontier orbitals of CTA are presented in Supple- CTA.
mentary Fig. 1, which shows that electron donor character is due to
the non-aromatic ring of cellulose. However, the electron with-
3.3.2. Effect of solution pH on dye removal
drawing nature is attributable to the aromatic ring and carboxyl
AO and ST removal is pH-dependent. Fig. 7d and e shows that
groups of the trimellitate moiety.
both AO and ST are positively charged in the pH range studied,
but ST may have one or two positive charges depending on the
3.2.2. Auramine-O (AO) solution pH. The PZC of the CTA adsorbent is 2.80. Thus, the
The neutral and monocation states of AO were considered (see adsorption of the cationic dyes AO and ST should be favored at
Fig. 3 for optimized structures), since dye removal was studied at pH > 2.8, where the CTA adsorbent has a net negative surface
multiple pH values. As can be seen in Fig. 3, the equilibrated struc- charge. Fig. 9 shows the removal of the dyes AO and ST as a func-
ture of the neutral AO is planar, while for the monocation state tion of solution pH from 2.13 to 7.14.
(AO+1) an approximately 30-degree dihedral angle formed between The removal of both dyes increased as the solution pH increased
the aromatic rings. for mono- and bi-component systems, and reached a plateau at pH
MED maps exhibited similar profiles for both species, showing values greater than 6 for AO and greater than 5 for ST, for the
higher electron densities localized on tertiary amines (red areas) monocomponent system. The adsorption of AO and ST increased
and lower densities in the central region (in blue), as shown in slightly until pH values greater than 7 for the bicomponent system.
Fig. 4. The frontier molecular orbitals HOMO and LUMO of AO At pH values greater than the PZC, the removal of both dyes is
are shown in Supplementary Fig. 2. expected to be controlled mainly by electrostatic interaction
between the negatively charged carboxylate groups on the CTA
3.2.3. Safranin-T (ST) adsorbent surface and the positively charged dye molecules. At
The +1 and +2 ionization states were considered for ST. ST+2 was low pH values (pH < 3), the CTA carboxylate groups were proto-
modeled by protonation of the primary amine. The electronic nated, substantially reducing removal of both dyes. At pH values
structure calculations carried out for the fully optimized ST+1 and closer to 2, the removal of AO was practically zero, while ST was
ST+2 showed that in both cases the phenyl rings are displaced per- still removed by the CTA adsorbent in both mono- and bi-
pendicularly to the condensed rings, as shown in Fig. 5. The MED component systems. Thus, at pH values lower than the PZC, the
maps have similar shapes for both cationic ST species, as shown removal of both dyes is expected to occur through hydrogen bond-
in Fig. 6. In these maps, it was observed that the amine groups in ing, dipole-dipole interactions, and p-p bonding (p stacking). In
the vicinity of the protonation sites have higher electron density addition, the adsorption capacities of AO and ST for the monocom-
(red color). The frontier molecular orbitals HOMO and LUMO of ponent system were higher than for the bicomponent system,
ST are shown in Supplementary Fig. 3. showing that in the pH range studied the adsorption of one dye
is suppressed by the presence another dye.
3.3. Adsorption studies Basic dyes such as AO and ST are primarily used to dye wool,
silk, acrylic fibers, and leather. In addition, exhaustion of dye baths
Fig. 7a–c shows suggested adsorption mechanisms for AO and is sensitive to pH change and therefore dye baths containing basic
ST removal from a single aqueous solution by the CTA adsorbent, dyes may have pH values between 4 and 5 depending on the type
as well as a desorption mechanism for a cationic dye (AO or ST) of fiber and its activation (e.g. presence of carboxylic or sulfonic
from the CTA adsorbent. acid groups [6]). Thus, in the present study, removal of dyes AO
and ST was studied at two pH values (i.e. 4.5 and 7.0) as the CTA
3.3.1. Effect of CTA dosage on dye removal adsorbent exhibits a good adsorption capacity at both pH values.
Fig. 8 shows the effect of CTA dosage on adsorption capacity and Removal of basic dyes at pH 4.5 is advantageous because the pH
removal percentage of AO. The removal percentage of AO increased of the dye effluent does not need to be changed, which decreases
as the CTA dosage was increased due to the increase in available the cost of the treatment. However, the removal of basic dyes at
adsorption sites. Contrarily, the amount of AO adsorbed per unit pH 7.0 can decrease the amount of adsorbent required to treat
weight of CTA (adsorption capacity) decreased with increasing the effluent, as the adsorption capacity of the CTA adsorbent at
pH 7.0 is greater than at pH 4.5.

3.3.3. Mono- and multi-component adsorption kinetics


Two critical parameters for designing a wastewater treatment
plant (WWTP) operating in batch or semicontinuous mode are
equilibrium time and adsorption rate [30]. Thus, to study the kinet-
ics of adsorption of dyes on CTA adsorbent, pseudo-first-order
(PFO) [31], pseudo-second-order (PSO) [32], Boyd [33], and intra-
particle diffusion (IPD) [34] models were used to model monocom-
ponent adsorption, while PFO, PSO, and a competitive kinetic
model of Corsel et al. [35] were used to model bicomponent
adsorption. These models were described in the first [10] and sec-
ond [11] parts of this series of studies.
Experimental data were modeled by nonlinear regression anal-
ysis (NRL) using Microcal OriginProÒ 2015 and MATLAB 2014
Fig. 2. Molecular electron density maps for the model of cellulose oligomer (Mathworks, Inc.) software. Data fit quality for the mono- and bi-
calculated using the B3LYP/6-31(2d,2p) level of theory. component experimental data were evaluated as described in the
580 F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590

Fig. 3. Fully optimized structures calculated for auramine-O in two states (a) neutral and (b) monocation, using the level of theory B3LYP/6-31(2d,2p). Nitrogen atoms are
blue, carbon are light brown, and hydrogen are white. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.)

Fig. 4. Molecular electron density maps calculated for the auramine-O in two states (a) neutral and (b) monocation AO+1, using the level of theory B3LYP/6-31(2d,2p). Red
color indicates high electron density, while blue indicates low electron density. (For interpretation of the references to colour in this figure legend, the reader is referred to the
web version of this article.)

Fig. 5. Fully optimized structure obtained for (a) ST+1 and (b) ST+2 using the level of theory B3LYP/6-31(2d,2p). Nitrogen atoms are blue, carbon light brown, and hydrogen are
white. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 6. Molecular electron density maps calculated for ST+1 (a) and ST+2 (b) using the level of theory B3LYP/6–31(2d,2p). Red indicates high electron density, while blue
indicates low electron density. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

first [10] and second [11] parts of this series of studies, able in Supplementary Tables 2 and 3. Fig. 10a–b shows plots of
respectively. qt over t for the monocomponent removal of AO and ST by the
CTA adsorbent, with the curves fit to experimental data using the
3.3.3.1. Monocomponent adsorption. Table 1 shows the parameters PFO and PSO kinetic models. As can be seen in Table 1, the PSO
estimated by modeling monocomponent experimental data with model best described AO and ST removal rate at both pH values,
the PFO, PSO, Boyd, and IPD models. Experimental data are avail- based on a higher coefficient of determination (R2), smaller
F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590 581

Fig. 7. Suggested adsorption mechanism of auramine-O (a) and safranin-T (b) on the CTA adsorbent. (c) Desorption mechanism of a cationic dye (AO or ST) from the CTA
adsorbent surface. Dyes structures as a function of the solution pH for (d) auramine-O and (e) safranin-T.

reduced chi-square (v2red), and a smaller residual sum of squares estimated by the PSO model for AO and ST were closest to qe,exp val-
(RSS). Furthermore, experimental adsorption capacities (qe,exp) were ues. In addition, the initial adsorption rate (h) for the removal of both
compared to estimated adsorption capacities (qe,est), qe,est values dyes by the CTA adsorbent was strongly affected by the solution pH.
582 F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590

Fig. 8. Effect of CTA dosage on AO adsorption capacity (s) and removal percentage (d) (C0 = 0.932 mmol L1, 130 rpm, 25 °C, and 24 h of agitation).

2 R  COOH þ dyeðaqÞ þ 2 H2 O
ðR  COOÞ2  dye þ 2 H3 Oþ

ð14Þ
Supplementary Fig. 4a–d present the intraparticle diffusion graphs
for monocomponent adsorption of AO and ST on the CTA adsorbent
at pH 4.5 and 7.0. These graphs demonstrate multilinear behavior
and the first straight lines for both dyes and pH values did not inter-
sect the origin, indicating that adsorption of dyes on the CTA adsor-
bent is initially controlled by film diffusion and then by intraparticle
diffusion until equilibrium is reached [36]. Considering that the
molecular sizes of AO and ST are 14.73 Å  4.92 Å and 11.36 Å 
8.10 Å, respectively, it is expected that adsorption of AO and ST
can occur in both micropores and mesopores, with adsorption in
micropores being slower than in mesopores due to micropores’
lower adsorbate-to-pore diameter ratio [37]. In addition, the calcu-
lated molecular volumes for neutral AO and AO+1 are 864.51 Å and
872.95 Å, while for ST+1 and ST+2 the values are 933.96 Å and
944.12 Å, respectively. The intraparticle diffusion graphs show
three stages for AO and ST adsorption at both pH values. The intra-
Fig. 9. Mono- and bi-component removal of auramine-O (AO) and safranin-T (ST) particle diffusion coefficients (ki) for the first and second stages of
by the CTA adsorbent as a function of the solution pH (25 °C, 130 rpm, 0.2 g L1 CTA, adsorption are presented in Table 1.
and 0.932 mmol L1 AO and ST for mono- and bi-component systems).
Supplementary Fig. 5a–d present the Boyd graphs for AO and ST
adsorption at pH 4.5 and 7.0. The slopes of the Boyd graphs (B)
The value of h for AO and ST increased 4 and 2.4 times, respectively,
were obtained from the linear portions of the graphs of Bt over t
as the solution pH increased from 4.5 to 7.0. Additionally, the equi-
and were used for calculating effective diffusion coefficients (Di)
librium adsorption capacities for AO and ST were higher at pH 7.0
[36] (Table 1). The Boyd graphs were linear in the initial adsorption
than at pH 4.5, indicating that the CTA adsorbent surface has a
period and did not intersect the origin. Values of B were close to
greater affinity towards AO and ST at higher pH values.
zero, suggesting that external mass transfer plays an important
As discussed above, the PSO kinetic model best described the
role in the adsorption of AO and ST, but lower values of ki,2 com-
removal of AO and ST by CTA adsorbent in the monocomponent
pared to ki,1 confirm that the rate-limiting step controlling the
system. Eqs. (13) and (14) represent two possible forms of interac-
adsorption of AO and ST is intraparticle diffusion.
tion of AO and ST onto CTA adsorbent at an adsorption site. Eq. (14)
best describe ST adsorption as the pKa of ST is 6.4 and at pH 7.0 a
3.3.3.2. Bicomponent adsorption. Table 2 shows the estimated
fraction of ST molecules still have two positive charges, while Eq.
kinetic parameters obtained by modeling bicomponent
(13) is most appropriate to describe AO adsorption. As suggested
experimental kinetic data with PFO, PSO, and Corsel models.
by the molecular electron density maps calculated for AO+1, ST+1,
Bicomponent experimental kinetic data are available in Supple-
and ST+2, two forms of interaction, that is electrostatic attraction
mentary Table 4. Supplementary Fig. 6a shows a graph of qt over
and van der Waals (dispersive forces, hydrogen bonding, and/or
t, with the curves fit to experimental data using the PFO and PSO
ion-dipole), may occur depending on the solution pH.
models for the AO-ST system at pH 7.0. Supplementary Fig. 6b
þ shows a graph of C overt t with the curves fit to the experimental
R  COOH þ dyeðaqÞ þ H2 O
R  COO  dye þ H3 Oþ ð13Þ
data using the Corsel model for bicomponent adsorption
F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590 583

Table 1
Results of modeling monocomponent experimental data for the removal of auramine-O and safranin-T (C0 = 0.932 mmol L1, 25 °C, 130 rpm and 0.2g L1 CTA).

Monocomponent systems
Parameters pH 4.5 pH 7.0
Auramine-O Safranin-T Auramine-O Safranin-T
te (min) 720 720 300 450
qe,exp (mmol g1) 2.721 ± 0.130 2.621 ± 0.190 3.111 ± 0.025 3.863 ± 0.045
Pseudo-first-order (PFO) kinetic model
qe,est (mmol g1) 2.222 ± 0.073 2.424 ± 0.095 3.037 ± 0.042 3.705 ± 0.059
k1 (min1) (1.48 ± 0.19)  102 (1.71 ± 0.00)  102 (4.03 ± 0.32)  102 (2.78 ± 0.21)  102
R2 0.9376 0.8548 0.9549 0.9673
v2red 0.0208 0.0373 0.0076 0.0110
RSS 0.2911 0.5589 0.1143 0.9673
Pseudo-second-order (PSO) kinetic model
qe,est (mmol g1) 2.449 ± 0.053 2.602 ± 0.070 3.226 ± 0.024 3.986 ± 0.041
k2 (g mmol1 min1) (8.53 ± 0.95)  103 (1.01 ± 0.15)  102 (1.96 ± 0.10)  102 (1.04 ± 0.07)  102
h (mmol g1 min1) (5.11 ± 0.58)  102 (2.04 ± 0.11)  101 (6.87 ± 1.02)  102 (1.66 ± 0.11)  102
R2 0.9805 0.9474 0.9904 0.9901
v2red 0.0065 0.0134 0.0016 0.0033
RSS 0.0911 0.2005 0.0241 0.4999
Intraparticle diffusion (IPD)
Step 1
ki,1 (mmol g1 min1/2) 0.215 ± 0.045 0.129 ± 0.021 0.297 ± 0.029 0.281 ± 0.022
C (mmol g1) -0.237 ± 0.259 0.460 ± 0.163 0.340 ± 0.181 0.522 ± 0.200
R2 0.9586 0.9065 0.9819 0.9751
Step 2
ki,2 (mmol g1 min1/2) 0.079 ± 0.009 0.037 ± 0.005 0.073 ± 0.009 0.018 ± 0.009
C (mmol g1) 0.776 ± 0.128 1.607 ± 0.09 2.044 ± 0.100 3.343 ± 0.163
R2 0.9697 0.9426 0.9731 0.4450
Boyd plot
B (min1) (6.06 ± 0.34)  103 (6.71 ± 0.74)  103 (1.15 ± 0.13)  102 (1.35 ± 0.05)  102
Di (m2 min1) 9.59  1012 9.77  1012 1.82  1011 2.14  1011
R2 0.9495 0.9195 0.9915 0.9293

Fig. 10. Adsorption kinetics of auramine-O (AO) and safranin-T (ST) on the CTA adsorbent in equimolar concentration (CAO = CST = 0.932 mmol L1) at 25 °C, 130 rpm, and 0.2
g L1. Monocomponent systems of AO (a) and ST (b) at pH 4.5 and 7.0 were modeled by pseudo-first- (PFO) and pseudo-second-order (PSO) kinetic models.

(Supplementary Table 5) for the AO-ST system at pH 7.0. Adsorp- monocomponent system. The model that best described bicompo-
tion of AO was suppressed by the presence of ST and likewise nent adsorption values for the systems AO-ST and ST-AO was the
adsorption of ST was suppressed by the presence of AO (Table 2). PFO kinetic model, which yielded higher R2, smaller v2red, and smal-
However, in bicomponent systems the decrease in qe for AO ler RSS. It should be noted that the adsorption mechanism was differ-
adsorption was greater than the decrease in qe for ST adsorption ent in monocomponent and bicomponent systems.
compared to qe values calculated in monocomponent systems. In As pointed out by Teodoro et al. [11], the PFO and PSO kinetic
addition, AO was adsorbed faster (4.5 times) on the CTA adsorbent models can only be used for comparing the adsorption rates and
than ST was. The value of te for AO adsorption decreased from 300 qe values of dyes in binary systems. These models cannot predict
to 240 min, while the value of te for ST remained constant (450 the influence of a dye i on the adsorption of a dye j, and vice versa.
min) when compared to values for AO and ST adsorption in the Therefore, a competitive kinetic model developed by Corsel et al.
584 F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590

Table 2 lower diffusion rates. Values of the unstirred layer (d) for bicompo-
Results of modeling bicomponent experimental data for the removal of auramine-O nent adsorption were small (Table 2).
and safranin-T (Ci = Cj = 0.932 mmol L1, 25 °C, 130 rpm and 0.2g L1 CTA).
Differences in adsorption kinetic parameters, that is rate con-
Bicomponent system (pH 7.0) stants, equilibrium adsorption capacities, and equilibrium times,
Parameters AO-ST ST-AO reveal that multicomponent adsorption of dyes requires further
te (min) 240 450
research, and the application of new adsorbent materials for treat-
qe.exp (mmol g1) 1.051 ± 0.029 3.258 ± 0.021 ment of effluents containing dyes must be studied in detail to pre-
Pseudo-first-order (PFO) kinetic model
dict adsorbent behavior for target dye adsorption.
qe.est (mmol g1) 1.057 ± 0.016 3.204 ± 0.226
k1 (min1) (7.22 ± 0.59)  102 (1.59 ± 0.41)  102 3.3.4. Mono- and multi-component adsorption isotherms
R2 0.9737 0.8617 Adsorption isotherms can elucidate the physical phenomenon
v2red 0.0035 0.1492
governing the removal of an adsorbate by an adsorbent in aqueous
RSS 0.0556 2.3878
phase at a constant temperature and pH [29]. In addition, adsorp-
Pseudo-second-order (PSO) kinetic model
tion isotherms are useful for predicting the maximum adsorption
qe.est (mmol g1) 1.108 ± 0.032 3.591 ± 0.318
k2 (g mmol1 min1) (8.89 ± 1.71)  102 (5.34 ± 2.25)  103 capacity of an adsorbent for one or more adsorbates and therefore
h (mmol g1 min1) (1.90 ± 0.21)  101 (6.88 ± 2.97)  102 are crucial to design of batch and semicontinuous adsorption sys-
R2 0.9213 0.8497 tems for WWTPs [39]. Five adsorption isotherm models were used
v2red 0.0104 0.1621
to describe the monocomponent removal of AO and ST by CTA
RSS 0.1664 2.5941
adsorbent: Langmuir [40], Sips [41], Fowler-Guggenheim (F-G)
Corsel kinetic model
[42], Hill-de Boer (H-B) [43,44], and Konda [45]. For a detailed
a 0.0129 0.0742
b 0.1322 0.7414
description of the first two isotherm models see the first part of
c 0.1451 0.8156 this series of studies [10]. In addition, IAST and RAST models were
ci - cj 0.6705 used to model bicomponent equilibrium data.
k+1 (L mmol1 min1) 64.35 181.89 The Fowler-Guggenheim (F-G) isotherm was developed by Fow-
k-1 (min1) 0.8714 5.0342
ler and Guggenheim [42]. It assumes localized adsorption and lat-
k+1/k-1 (L mmol1) 73.8467 36.1309
d (m) 1.829  109 eral interaction between adsorbed molecules on the surface of an
Di (m2 min1)b 1.82  1011 2.14  1011 adsorbent. The F-G isotherm is given by Eq. (15).
Cmax (mmol m2) 0.321 1.024
qt (mmol g1)a 1.026 3.271 h
R2 K FG C e ¼ expðchÞ ð15Þ
0.8364 0.8962 1h
v2red 0.0061 0.0094
RSS 0.0216 0.1014 where KFG (L mmol1) is the Fowler-Guggenheim constant, h is frac-
a
Calculated from a specific surface area of 3.196 ± 0.217 m2 g1 (Teodoro et al. tional surface coverage (qe Q1 max), and c (kJ mol
1
) is a constant
[10]). related to the energy involved in the interaction between adsorbed
b
Converted values from the data reported by Teodoro et al. [10]. molecules. If the interaction between adsorbed molecules is attrac-
tive, the c > 0 and the heat of adsorption will increase with increas-
ing surface coverage. On the contrary, if the interaction between
[35] for protein adsorption was adopted to evaluate the kinetics of
adsorbed molecules is repulsive, c < 0 and the heat of adsorption
the bicomponent systems AO-ST and ST-AO. The kinetic model of
will decrease with increasing the surface loading. When there is
Corsel et al. [35] is based on a generalized diffusion equation,
no lateral interaction between adsorbed molecules c = 0 and the
allowing it to include the effects of surrounding bulk medium
F-G isotherm reduces to the Langmuir isotherm [46].
and an unstirred layer. The association constant (K = k+1/k-1), which
The Hill-de Boer (H-B) isotherm was developed by Hill [43] and
is the ratio between the intrinsic adsorption (k+1) and desorption
de Boer [44]. It assumes mobile adsorption and lateral interaction
(k-1) rate constants, was higher for AO than for ST, suggesting that
between adsorbed molecules. The H-B isotherm is given by Eq.
the adsorption sites of the CTA adsorbent had a higher affinity for
(16).
AO than for ST (Table 2).
 
The interaction constants a and b can give relevant information h h
on possible lateral interactions between adsorbed solutes in a K HB C e ¼ exp  ch ð16Þ
1h 1h
bicomponent system. Smaller positive values of cooperative con-
stants (a) (Table 2) imply that dye molecules do not tend to adsorb where KHB (L mmol1) is the H-B constant. When c is positive, there
near other adsorbed dye molecules. This inhibition is probably due is an attractive force between adsorbed molecules, while a negative
to a repulsive potential around adsorbed dye molecules or steric value means there is repulsion between adsorbed molecules. When
hindrance of the adsorbed dye molecules, which may hinder dye c = 0, the H-B isotherm is reduced to the Volmer isotherm that
molecules’ adsorption on adjacent sites [38]. Therefore, the higher assumes mobile adsorption but no lateral interaction between
the value of cooperative constants (a), the greater the decrease in adsorbed molecules [46].
the adsorption rate function as the CTA adsorbent surface becomes The Konda isotherm was developed by Konda et al. [45] to
saturated. On the other hand, the higher the value of cooperative describe single- and multi-step adsorption isotherms of organic
constants (b), the greater the increase in the desorption rate func- pesticides on soil. This isotherm is derived from the Langmuir iso-
tion as the CTA adsorbent surface becomes saturated. therm with the additional assumption that sterically or energeti-
If dkonCmax < < D, diffusion has no influence on adsorption cally heterogeneous adsorption sites are present on the
kinetics [35]. For AO and ST adsorption, values of dkonCmax were adsorbent surface. It also assumes potential interactions between
3.92  1011 and 3.41  1010, respectively. Thus, as pointed out adsorbed molecules, and that each step on the isotherm represents
for monocomponent adsorption of AO and ST, diffusion played an a different type of adsorption mechanism [45]. The Konda model is
important role in the bicomponent adsorption process of AO and given by Eq. (17).
ST on the CTA adsorbent. The thickness of the unstirred layer of
Q 1 b1 C e Q b2 ½ðC e  c2 Þ þ jC e  c2 j
water adjacent to CTA adsorbent particles can play an important qe ¼ þ 2 ð17Þ
1 þ b1 C e 2 þ b2 ½ðC e  c2 Þ þ jC e  c2 j
role in the adsorption process, because thicker layers result in
F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590 585

where Qi (mmol g1) is the adsorption capacity for the ith adsorp- numbers (CN) [54] of the CTA adsorbent for AO and ST adsorption
tion step, bi (L mmol1) is the adsorption equilibrium constant, at pH 4.5 were greater than unity (i.e. 2.40 and 1.85 for AO and ST,
and ci (mmol L1) is the limit concentration. The limit concentration respectively), indicating that more than one carboxylic group is
for the first adsorption step is zero. The maximum adsorption adsorbing to a dye molecule.
P
capacity is defined by Qmax = Qi. The adsorption isotherm of ST on the CTA adsorbent at pH 7.0
The ideal adsorbed solution theory (IAST) was proposed by (Fig. 11b) was type I, which was well described by the Langmuir
Myers and Prausnitz [47] for adsorption of a gas mixture and and Sips models. The value of the parameter n (0.67) of the Sips
was extended to the liquid phase by Radke and Prausnitz [48]. model does not allow reducing the Sips model to the Langmuir
The IAST model provides a thermodynamically consistent and model, although the values of isotherm parameters Qmax and b
practical method for predicting bicomponent adsorption isotherms for the Sips model were similar to those obtained for the Langmuir
using only monocomponent isotherm data [13]. The RAST model model. The parameter n of the Sips model indicates that ST adsorp-
was proposed by Costa et al. [49] and is an extension of the IAST tion on the CTA adsorbent is heterogeneous, which is more consis-
model proposed by Myers [50] and Talu and Zwiebel [51] for mul- tent for this adsorption system than the assumptions of the
ticomponent adsorption of non-ideal gas mixtures. The RAST Langmuir model. On the other hand, the adsorption isotherm of
model is more appropriate than the IAST model to describe adsorp- AO at pH 7.0 (Fig. 11a) was type V or S-shaped isotherm. This type
tion systems that exhibit synergistic or antagonistic behavior. The of isotherm occurs when there is lateral interaction between
RAST model is a derivation of the IAST model. However, it assumes adsorbed molecules. As the Langmuir model does not assume
that the adsorbed phase is deviates from ideality [11]. adsorbate-adsorbate interaction, the Hill-de Boer and Fowler-
Multicomponent adsorption equilibria using the IAST and RAST Guggenheim models were used to model the adsorption isotherm
models is predicted by two variables for IAST (i.e. the solid-phase of AO on the CTA at pH 7.0. As seen in Table 3, values of the con-
mole fractions and the reduced spreading pressure), and six vari- stant c for both isotherm models were positive, suggesting that
ables for RAST (i.e. the solid-phase mole fractions, the reduced there are attractive interactions between adsorbates on the CTA
spreading pressure, and adjustable parameters of the activity coef- surface. Both models fit the experimental adsorption data well.
ficient model). The spreading pressure-dependent Wilson equation However, comparing the experimental adsorption capacity
was used for determination of the activity coefficients (c, Ʌ12, and (Qmax,exp) with the estimated adsorption capacity (Qmax,est), the F-
Ʌ21). The Sips isotherm was used to model monocomponent G model better represented the adsorption process than the H-B
adsorption equilibria to obtain the model parameters Qmax, b, and model. On the other hand, the H-B model resulted in smaller values
n. The theoretical approach for the IAST and RAST models and of v2red and RSS compared to the F-G model. Therefore, analysis of fit
the solution methods used for both models is presented in detail quality for the F-G and H-B models does not allow determination of
in the second part of this series of studies [11], in the Supplemen- whether mobile adsorption occurred on the CTA surface. In addition,
tary Material, and in the studies of Erto [52,53]. As reported by Erto CN values for AO and ST adsorption on the CTA adsorbent were 1.25
et al. [53], linear regression analysis using the modified Wilson and 1.58, respectively. These values indicate that CN decreased as the
equation parameters ci and cj (see Eqs. (36) and (37) reported by solution pH increased, suggesting a change in the mechanism by
Teodoro et al. [11]) for the entire experimental data did not result which AO and ST molecules are adsorbed on the CTA surface. That
in a good linear fit. Thus, the approximation suggested by Erto et al. is, fewer adsorption sites are needed for adsorption of AO and ST,
[53] (see Eqs. (17)–(21) described by Erto et al. [53]) was applied in and therefore, the adsorption capacity at pH 7.0 increased.
the present study to calculate the value of cj from the value of ci,
which was obtained by regression analysis of the modified Wilson 3.3.4.2. Bicomponent adsorption. Fig. 11c shows a plot of qe against
equation (Eq. (36) described in the study of Teodoro et al. [11]). Ce for the bicomponent system AO-ST (Supplementary Table 8)
with curves fit to the experimental data using the IAST and RAST
3.3.4.1. Monocomponent adsorption. Table 3 shows the parameters models supplied with monocomponent Sips isotherm parameters
estimated by modeling the monocomponent experimental data Qmax, b, and n (Table 3) obtained by fitting both monocomponent
(Supplementary Tables 6 and 7) with the Langmuir, Sips, Fowler- experimental equilibrium data to this model. Table 4 shows the
Guggenheim, Hill-de Boer, and Konda isotherm models through values of the error functions used in the minimization of the resid-
NLR analysis using OriginProÒ 2015 and MATLAB 2014 (Math- ual sum of squares (RSS) between the experimental initial liquid
Works, Inc.) software. NLR analysis was as described in Sec- phase solute concentrations and those predicted by the IAST and
tion 3.3.3. Fig. 11a and b shows plots of qe against Ce for the RAST models, as well as the coefficient of determination (R2) and
monocomponent adsorption of AO and ST on the CTA adsorbent, reduced chi-square (v2red). As can be seen, the quality of the predic-
with isotherm curves fit to experimental data. tion of bicomponent adsorption equilibrium data for ST using the
Notably, the monocomponent adsorption isotherms of AO and IAST and RAST models was very good (R2 > 0.98) while for AO it
ST exhibited different shapes. The adsorption isotherms of AO was satisfactory (R2 > 0.81). Supplementary Fig. 7 shows graphs of
and ST at pH 4.5 can be classified as type IV isotherms because they qe,est against qe,exp, which are used to analyze if the experimental
exhibited more than one adsorption step, indicating the inhomo- data can be predicted by the model. Both models described the bin-
geneity of the CTA adsorbent surface. It seems that most of the ary system AO-ST well.
adsorption sites on the CTA adsorbent surface are saturated at a The data presented in Table 3 and Fig. 11a–c show that the
relatively low equilibrium concentration (limit concentration), adsorption capacity of the CTA adsorbent for both AO and ST was
and interactions between adsorbed molecules or conformational greater in the monocomponent systems than the bicomponent sys-
changes of adsorbed molecules on the CTA adsorbent surface tem. This indicates the occurrence of antagonistic interactions
may occur, allowing the adsorption of more AO and ST molecules. between dyes, which decreased their adsorption capacity in the
The Konda model described the adsorption steps of AO and ST on bicomponent system. The adsorption capacity of AO was more
CTA very well and provided good estimates of Qmax (Fig. 11 and greatly affected than that of ST.
Table 3). The binding energy (b1) of ST with adsorption sites of The plot of gamma of the component i against gamma of the
the CTA adsorbent was greater than AO for the first adsorption component j was nonlinear (Supplementary Fig. 8), so the activity
step, while the same behavior was noticed for the binding energy coefficient of the component j was calculated using Eqs. (17)–(21)
(b2) of the second adsorption step, indicating that ST was bound reported by Erto et al. [53]. The adjustable coefficients a and b
more strongly than AO on CTA adsorption sites. The coordination obtained by the approximation suggested by Erto et al. [53] are
586 F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590

Table 3
Results of modeling the monocomponent equilibrium adsorption data of auramine-O and safranin-T with isotherm models (25 °C, 130 rpm and 0.2 g L1 CTA).

Monocomponent adsorption systems


Isotherm model Parameters pH 4.5 pH 7.0
Auramine-O Safranin-T Auramine-O Safranin-T
Experimental data Qmax.exp (mmol g1) 2.628 ± 0.052 3.443 ± 0.059 5.172 ± 0.006 4.019 ± 0.147
te (min) 720 720 300 450
l (mol L1) 0.0775 0.0774 – 0.1507
ce 0.7744a 0.7745a – 0.7594b
Langmuir Qmax.est (mmol g1) 2.531 ± 0.078 3.262 ± 0.164 – 4.314 ± 0.120
b (L mmol1) 17.80 ± 2.02 23.57 ± 4.65 34.31 ± 3.69
R2 0.9417 0.8180 0.9277
v2red 0.0097 0.0487 0.0114
RSS 0.1352 0.8772 0.1482
Sips Qmax.est (mmol g1) – – 5.055 ± 0.107 4.074 ± 0.143
b (L mmol1) 3.03 ± 0.04 32.64 ± 2.37
n 0.17 ± 0.01 0.67 ± 0.14
1/n 5.768 ± 0.258 1.499 ± 0.307
R2 0.9942 0.9432
v2red 0.0114 0.0101
RSS 0.2279 0.1114
Hill-de Boer KHB (L mmol1) – – 0.318 ± 0.050 –
Qmax,est (mmol g1) 7.386 ± 0.701
c (kJ mol1) 6.223 ± 0.533
R2 0.9856
v2red 0.1246
RSS 1.1491
Fowler-Guggenheim KFG (L mmol1) – – 0.484 ± 0.077 –
Qmax,est (mmol g1) 5.443 ± 0.504
c (kJ mol1) 3.616 ± 0.361
R2 0.9860
v2red 0.1620
RSS 1.2406
Konda Q1 (mmol g1) 2.200 2.600 – –
Q2 (mmol g1) 0.641 1.091
Qmax.est (mmol g1) 2.841 3.691
b1 (L mmol1) 26.575 47.9774
b2 (L mmol1) 7.983 17.5653
c2 (mmol L1) 0.482 0.320
R2 0.9883 0.9710
v2red 0.0261 0.1902
RSS 0.0453 0.3286
a
Calculated using Guntelberg approximation.
b
Calculated using Davies approximation.

presented in Table 5. The plots of the experimental activity coeffi- coverage. Fig. 12 shows DadsH values as a function of the amount
cients against experimental reduced spreading pressures are of the dyes adsorbed on CTA at equilibrium at pH 4.5 and 7.0.
shown in Supplementary Fig. 9. This plot presents an increasing DadsH values were all negative in the qe range evaluated,
trend for the activity coefficient for one component when the indicating that adsorption phenomenon is exothermic. DadsH val-
spreading pressure is increased and an opposite behavior for ues varied from 18.83 ± 0.64 kJ mol1 to 5.60 ± 0.02 kJ mol1,
another component. suggesting that physisorption governs the adsorption process.
Energy dispersive X-ray (EDX) spectroscopy (SEM-EDX) was For the adsorption of ST, the profile of the curves of DadsH
used to map the distribution of the adsorbed AO and ST along against qe as well as the magnitude of DadsH depended strongly
the CTA surface for the mono- and bi-component systems. Results on the solution pH, confirming the large contribution of electro-
are discussed and presented in the Supplementary Material. static interactions to dye adsorption. For this dye, as the qe value
increased, the DadsH values became less exothermic at pH 4.5
and more exothermic at pH 7.0. DadsH values at pH 4.5 were always
3.4. Thermodynamics of adsorption more negative than at pH 7.0. To explain these results, DadsH was
broken into three terms associated with three independent
3.4.1. Isothermal titration calorimetry subprocesses occurring simultaneously during dye adsorption, as
Isothermal titration calorimetry (ITC) is a sensitive technique presented in Eq. (18).
used to measure the enthalpy change associated with a specific
thermodynamic process occurring within a calorimetric cell when Dads H ¼ Dads Hdes þ Dads HdyeCTA þ Dads Hdyedye ð18Þ
the system composition is varied by titration. In the first part of
this series of studies, ITC was successfully used to obtain the where DadsHdes is the enthalpy change associated with the desolva-
enthalpy change associated with the adsorption of metal ions on tion of both cationic dyes and adsorption sites on the CTA surface,
the CTA adsorbent [10]. In the same way, ITC was used in the pre- DadsHdye-CTA is the enthalpy change for the formation of the interac-
sent study to determine the adsorption enthalpy change (DadsH) of tion between the dye and CTA, and DadsHdye-dye is the enthalpy
AO and ST on the CTA as a function of the adsorbent surface change related to the dye-dye interaction on the adsorbent surface.
F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590 587

Fig. 11. Adsorption isotherms of auramine-O (AO) and safranin-T (ST) on the CTA adsorbent in equimolar concentration (Ci = Cj) at 25 °C, 130 rpm, and 0.2 g L1 CTA.
Monocomponent adsorption systems of (a) AO and (b) ST at pH 4.5 and 7.0. Bicomponent adsorption system of (c) AO-ST at pH 7.0.

Table 4
Results of modeling the equilibrium adsorption data with the IAST and RAST model for bicomponent system AO-ST (25 °C, 130 rpm, pH 7.0 and 0.2 g L1 CTA with equimolar dye
concentrations (Ci = Cj)).

Adsorption system Dye Model RSS R2 v2red


AO-ST AO IAST 0.3368 0.8526 2.410  102
RAST 0.3981 0.8171 2.345  102
ST IAST 2.4278 0.9943 5.754  102
RAST 1.8076 0.9827 4.559  102

Table 5
Values of adjustable parameters of the Wilson activity coefficient model for bicomponent system AO-ST (i = 1 and j = 2) and of the model proposed by Erto et al. [58].

Adsorption system (i-j) c K12 K21 a1 b2 R2i R2j RSS


AO-ST 53.8366 0.2660 2.0280 – – 0.9975 0.8895 1.005
– – – 1.5508 1.7462 0.9230 –
1,2
Values obtained by regression analysis of Eq. (17) according to the approximation suggested by Erto et al. [53].

The dye-dye interaction in solution was disregarded because the pH and low qe values (i.e. the condition under which DadsH was
dye equilibrium concentrations in the calorimetric experiments obtained), the solvation shells of both dyes and adsorption sites
were very low (in the range from 107 to 105 mol L1). At constant should remain unchanged, and the term DadsHdes is constant as
588 F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590

Table 6
Thermodynamic parameters of monocomponent adsorption of auramine-O (AO) and
safranin-T (ST) on the CTA adsorbent at 25 °C (298 K).

Thermodynamic AO ST ST
parameters
pH 4.5 pH 7.0
DadsG° (kJ mol1) 25.88 ± 0.00 27.35 ± 0.00 26.45 ± 0.00
DadsH° (kJ mol1) 7.24 ± 0.09 19.10 ± 0.79 1.66 ± 0.39
TDadsS° (kJ mol1) 18.65 ± 0.22 8.26 ± 0.34 24.79 ± 5.84

parameters (DadsG°, DadsH°, and DadsS°) were determined (Table 6).


Values of DadsG° were determined using Eq. (11) while values of
DadsH° were obtained by extrapolating the curve of DadsH against
Ce to Ce equals zero. The entropic contribution, TDadsS°, was deter-
mined from the fundamental relationship given by Eq. (19).
Dads G ¼ Dads H  T Dads S ð19Þ

Values of DadsG° were negative for all conditions evaluated,


Fig. 12. Adsorption enthalpy change as a function of the amount adsorbed at showing that the dyes preferentially concentrate on the CTA surface
equilibrium for AO and ST on the CTA surface at 25.0 °C: (s) AO-pH 4.5, (h) ST- under standard conditions. Moreover, values of DadsH° and
pH 4.5, and (j) ST-pH 7.0. TDadsS° were always negative and positive, respectively, indicating
that the adsorption process under standard conditions is enthalpi-
cally and entropically favored. The entropy increase is attributed
the qe value increases. Thus, the change in DadsH when the surface
to the water molecules released from the solvation shell of both
coverage increases is mainly attributed to the subprocesses associ-
dyes and CTA, mainly due to the formation of dye-CTA hydrophobic
ated with the formation of the dye-CTA and dye-dye interactions.
interactions.
Adsorption of the dyes on the CTA surface is mainly controlled
by electrostatic interactions between the negatively charged car-
3.5. Desorption
boxylate groups on the CTA surface and the positively charged
dye molecules. Consequently, the greater the force of the
3.5.1. Type of desorption agent
dye-CTA electrostatic attraction, the greater the magnitude of
Two types of desorption solutions were used in the experiments,
DadsHdye-CTA (always negative). ST has two positive charges at pH
HNO3 and HCl, at concentrations of 0.01 and 0.1 mol L1. Table 7
4.5, while ST mostly has one positive charge at pH 7.0 (pKa of ST
shows that aqueous 0.1 mol L–1 HCl was the best desorption agent
is 6.4). Thus, the dye-CTA electrostatic attraction is more intense
for ST, while aqueous 0.01 mol L–1 HNO3 was the best desorption
at pH 4.5, making the term DadsHdye-CTA more negative and there-
agent for AO. However, the Edes of aqueous 0.1 mol L–1 HCl was
fore DadsH more exothermic at this pH.
not much better than the Edes of aqueous 0.01 mol L–1 HCl for ST
When the surface coverage increases at pH 4.5, the distance
desorption. Similarly, the Edes of aqueous 0.01 mol L–1 HNO3 was
between positively charged ST molecules on the CTA surface
not much better to the Edes of aqueous 0.01 mol L–1 HCl for AO
decreases, increasing the dye-dye electrostatic repulsion on the
desorption. Therefore, for practical purposes, the same desorption
adsorbent surface. Thus, the term DadsHdye-dye becomes more pos-
solution can be used for both dyes. Aqueous 0.01 mol L–1 HCl solu-
itive and the values of DadsH increase as qe increases at this pH.
tion was chosen for the continuation of desorption experiments as a
When the pH is changed from 4.5 to 7.0, dye-dye electrostatic
function of time.
repulsion on the CTA surface decreases due to deprotonation of
ST, making the term DadsHdye-dye less positive at pH 7.0 than at
4.5. Nonetheless, an increase in DadsH as qe increases would also 3.5.2. Effect of contact time on desorption
be expected at pH 7.0. However, DadsH decreased when the surface After determining the best desorption agent and solution con-
coverage increased, suggesting that different conformations of car- centration for desorption of AO and ST from CTA adsorbent, the
boxylic ligands and/or adsorbed ST molecules on the CTA may have effect of contact time on Edes was evaluated. Table 7 shows that Edes
occurred. increased as desorption time increased from 1 to 3 h for both dyes.
For adsorption of AO on CTA at pH 4.5, the values of DadsH were However, there was no significant increase in Edes after 3 h. As
less negative than those for ST at the same pH, which was expected desorption implies an additional process cost, shorter times are
due to the higher positive net charge of ST (i.e. DadsHdye-CTA is more preferable. Therefore 3 h was chosen as the optimal desorption
negative for ST than for AO). Moreover, values of DadsH for AO time for both dyes. Incomplete desorption of AO and ST from the
adsorption at pH 4.5 were almost independent of qe, suggesting CTA surface suggests that ion-exchange is not the sole mechanism
that the energies involved in the dye-dye interaction and the con- controlling the desorption of AO and ST, and therefore, these dyes
formational changes on the CTA surface may compensate for each also form strong interactions with CTA that cannot be broken by
other. Values of DadsH were not obtained for AO at pH 7.0 because the desorption solutions tested.
in the AO concentration range evaluated (limited by the dye solu-
bility), dye adsorption was very low (Fig. 12) and the signal mea- 3.6. Re-adsorption
sured in the calorimetry experiment was of the same magnitude
as equipment noise. Re-adsorption efficiency (Ere-ads) is a very important process
variable for an adsorbent material and will define if the adsorbent
material has the possibility of being used in various adsorption
3.4.2. Thermodynamic parameters of adsorption cycles, and therefore, determine if its use is economic feasible.
To complete the thermodynamic analyses of the adsorption Table 7 shows that Ere-ads for AO and ST were not 100% because
process of AO and ST on CTA, standard adsorption thermodynamic both AO and ST were not fully desorbed from the CTA adsorbent.
F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590 589

Table 7
Desorption efficiency (Edes) as a function of desorption agent, desorption solution concentration and time and re-adsorption efficiency (Ere-ads) for auramine-O (AO) and safranin-T
(ST).

Type of desorption agent HNO3 HCl


Dye Concentration (mol L1) 0.01 0.1 0.01 0.1
Auramine-O Edes (%)a 27.34 ± 0.03 4.51 ± 0.08 21.57 ± 0.13 3.95 ± 0.03
Safranin-T 46.26 ± 0.10 32.25 ± 0.05 74.83 ± 0.16 78.35 ± 0.16
Effect of contact time on desorption efficiency
Dye Time (h) 1 3 6 12 24
Auramine-O Edes (%)b 38.23 ± 0.59 44.61 ± 0.44 44.69 ± 0.53 45.54 ± 0.41 47.54 ± 0.33
Safranin-T 65.58 ± 0.46 78.87 ± 0.49 81.45 ± 0.51 82.64 ± 0.65 81.12 ± 0.18
Re-adsorption efficiency
Dye QT,max (mmol g1) Qe,re-ads (mmol g1) Ere-ads (%)
Auramine-O 2.67 ± 0.01 2.08 ± 0.06 77.90 ± 0.15
Safranin-T 3.64 ± 0.01 2.21 ± 0.01 60.71 ± 0.05
a
Volume of desorption solution was 20.0 mL.
b
Volume of desorption solution was 100.0 mL.

However, Ere-ads for AO was higher than Ere-des for ST, suggesting viding the equipment and technical support for experiments
that the surface of the CTA adsorbent had a higher affinity for AO involving electron microscopy and SEM-EDX analyses. This work
molecules after desorption. In addition, Ere-ads for ST was smaller is a collaboration research project of members of the Rede Mineira
than for AO, indicating a loss of adsorption capacity of ST by CTA de Química (RQ-MG) supported by FAPEMIG (Project: CEX-RED-
after desorption. Similar observations were noticed by Ferreira 00010-14).
et al. [29] for desorption and re-adsorption of crystal violet (CV)
from sugarcane bagasse modified with Meldrum’s acid. Therefore,
desorption and re-adsorption experiments demonstrated that the Appendix A. Supplementary material
CTA adsorbent can be reused, but rigorous control is necessary to
ensure that dye effluent concentrations are within the limits of dis- Supplementary data associated with this article can be found, in
charge due to the partial desorption of dyes and the unpredictable the online version, at https://doi.org/10.1016/j.jcis.2017.10.083.
re-adsorption behavior of the CTA adsorbent. The mass balance for
desorption and re-adsorption experiments is presented in Supple-
References
mentary Table 9.
[1] V.K. Gupta, R. Kumar, A. Nayak, T.A. Saleh, M.A. Barakat, Adsorptive removal of
dyes from aqueous solution onto carbon nanotubes: a review, Adv. Colloid
4. Conclusions
Interface Sci. 193–194 (2013) 24–34.
[2] V.K. Gupta, R. Jain, S. Varshney, V.K. Saini, Removal of Reactofix Navy Blue 2
Cellulose modified with trimellitic anhydride (CTA) was very GFN from aqueous solutions using adsorption techniques, J. Colloid Interface
efficient to remove AO and ST from aqueous solutions, demonstrat- Sci. 307 (2007) 326–332.
[3] M.T. Yagub, T.K. Sen, S. Afroze, H.M. Ang, Dye and its removal from aqueous
ing excellent monocomponent adsorption capacities at both pH 4.5 solution by adsorption: a review, Adv. Colloid Interface Sci. 209 (2014) 172–
and 7.0. In bicomponent systems, the presence of ST affected the 184.
adsorption of AO more than the presence of AO affected the [4] M. Vakili, M. Rafatullah, B. Salamatinia, A.Z. Abdullah, M.H. Ibrahim, K.B. Tan, Z.
Gholami, P. Amouzgar, Application of chitosan and its derivatives as
adsorption of ST on the CTA adsorbent. Despite the complexity of adsorbents for dye removal from water and wastewater: a review,
these systems, both the IAST and RAST models satisfactory fit the Carbohydr. Polym. 113 (2014) 115–130.
bicomponent equilibrium data. Desorption of dyes from the CTA [5] V.K. Gupta, Suhas, Application of low-cost adsorbents for dye removal – a
review, J. Environ. Manage. 90 (2009) 2313–2342.
adsorbent was incomplete. However, CTA adsorbed more dye in a [6] K. Venkataraman, The Chemistry Of Synthetic Dyes, Academic Press, New York,
second adsorption cycle, showing the reliability of reuse of CTA 1971.
adsorbent. The use of isothermal titration calorimetry allowed [7] R.W. Sabnis, Handbook of Biological Dyes and Stains: Synthesis and Industrial
Applications, John Wiley & Sons, New Jersey, 2010.
evaluation of how the magnitude of the enthalpy of adsorption [8] N. Terinte, B.M.K. Manda, J. Taylor, K.C. Schuster, M.K. Patel, Environmental
depended on surface coverage, showing that the adsorption sites assessment of coloured fabrics and opportunities for value creation: spin-
on the CTA surface are distinct and that dye-dye interactions affect dyeing versus conventional dyeing of modal fabrics, J. Clean. Prod. 72 (2014)
127–138.
adsorption energies. The possibility of using the CTA adsorbent to
[9] M.A.M. Salleh, D.K. Mahmoud, W.A.W.A. Karim, A. Idris, Cationic and anionic
remove dyes from liquid phase at different pH values shows how dye adsorption by agricultural solid wastes: a comprehensive review,
versatile this material is, which could reduce costs and improve Desalination 280 (2011) 1–13.
efficiency of an effluent treatment process. [10] F.S. Teodoro, S.N.D.C. Ramos, M.M.C. Elias, A.B. Mageste, G.M.D. Ferreira, L.H.M.
da Silva, L.F. Gil, L.V.A. Gurgel, Synthesis and application of a new carboxylated
cellulose derivative. Part I: removal of Co2+, Cu2+ and Ni2+ from
Acknowledgements monocomponent spiked aqueous solution, J. Colloid Interface Sci. 483 (2016)
185–200.
[11] F.S. Teodoro, O.F.H. Adarme, L.F. Gil, L.V.A. Gurgel, Synthesis and application of
The authors are grateful to Universidade Federal de Ouro Preto a new carboxylated cellulose derivative. Part II: removal of Co2+, Cu2+ and Ni2+
(UFOP grant number 23109.003209/2016-98) and Fundação de from bicomponent spiked aqueous solution, J. Colloid Interface Sci. 487 (2017)
266–280.
Amparo à Pesquisa do Estado de Minas Gerais (FAPEMIG grant
[12] K.A.G. Gusmão, L.V.A. Gurgel, T.M.S. Melo, L.F. Gil, Adsorption studies of
numbers APQ-01945-13 and APQ-01287-15) for funding this methylene blue and gentian violet on sugarcane bagasse modified with EDTA
research. The authors are also grateful to Coordenação de Aper- dianhydride (EDTAD) in aqueous solutions: kinetic and equilibrium aspects, J.
feiçoamento de Pessoal de Nível Superior (CAPES). The authors Environ. Manage. 118 (2013) 135–143.
[13] O.S. Chan, W.H. Cheung, G. McKay, Single and multicomponent acid dye
would like to acknowledge the Nanolab Electronic Microscopy Lab- adsorption equilibrium studies on tyre demineralised activated carbon, Chem.
oratory, at the Redemat, Escola de Minas, UFOP, MG, Brazil, for pro- Eng. J. 191 (2012) 162–170.
590 F.S. Teodoro et al. / Journal of Colloid and Interface Science 512 (2018) 575–590

[14] Y. Liu, Is the free energy change of adsorption correctly calculated?, J Chem. [35] J.W. Corsel, G.M. Willems, J.M.M. Kop, P.A. Cuypers, W.T. Hermens, The role of
Eng. Data 54 (2009) 1981–1985. intrinsic binding rate and transport rate in the adsorption of prothrombin,
[15] E. Guntelberg, Untersuchungen uber ioneninteraction, Z. Physik. Chem. 123 albumin, and fibrinogen to phospholipid bilayers, J. Colloid Interface Sci. 111
(1926) 199–247. (1986) 544–554.
[16] C.W. Davies, Ion Association, Butterworths, London, UK, 1962, pp. 37–53. [36] B.H. Hameed, M.I. El-Khaiary, Malachite green adsorption by rattan sawdust:
[17] P.S. Ghosal, A.K. Gupta, Determination of thermodynamic parameters from isotherm, kinetic and mechanism modeling, J. Hazard. Mater. 159 (2008) 574–
Langmuir isotherm constant-revisited, J. Mol. Liq. 225 (2017) 137–146. 579.
[18] F. Neese, The ORCA program system, Wiley Interdisciplinary Rev.: Comput. [37] X.-Y. Yang, B. Al-Duri, Application of branched pore diffusion model in the
Mol. Sci. 2 (2012) 73–78. adsorption of reactive dyes on activated carbon, Chem. Eng. J. 83 (2001) 15–23.
[19] P. Hohenberg, W. Kohn, Inhomogeneous electron gas, Phys. Rev. 136 (1964) [38] V. Hlady, Adsorption kinetics of low density lipoprotein onto a hydrophobic-
B864–B871. hydrophilic gradient surface, in: G. Pifat-Mrzljak (Ed.), Supramolecular
[20] W. Kohn, L.J. Sham, Self-consistent equations including exchange and Structure and Function 7, Springer, Rovinj, Croatia, 2001, pp. 45–61.
correlation effects, Phys. Rev. 140 (1965) A1133–A1138. [39] K.Y. Foo, B.H. Hameed, Insights into the modeling of adsorption isotherm
[21] A.D. Becke, Density-functional thermochemistry. III. The role of exact systems, Chem. Eng. J. 156 (2010) 2–10.
exchange, J. Chem. Phys. 98 (1993) 5648–5652. [40] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and
[22] C. Lee, W. Yang, R.G. Parr, Development of the Colle-Salvetti correlation-energy platinum, J. Am. Chem. Soc. 40 (1918) 1361–1403.
formula into a functional of the electron density, Phys. Rev. B 37 (1988) 785– [41] R. Sips, On the structure of a catalyst surface, J. Chem. Phys. 16 (1948) 490–
789. 495.
[23] A.J. Cohen, P. Mori-Sánchez, W. Yang, Challenges for density functional theory, [42] R.H. Fowler, E.A. Guggenheim, Statistical Thermodynamics, Cambridge
Chem. Rev. 112 (2012) 289–320. University Press, London, 1939.
[24] L.C. Cardoso, R.M.L. Savedra, M.M. Silva, G.R. Ferreira, R.F. Bianchi, M.F. [43] T.L. Hill, Statistical mechanics of multimolecular adsorption II. Localized and
Siqueira, Effect of blue light on the electronic and structural properties of mobile adsorption and absorption, J. Chem. Phys. 14 (1946) 441–453.
Bilirubin isomers: insights into the photoisomerization and photooxidation [44] J.H. de Boer, The Dynamical Character of Adsorption, Oxford University Press,
processes, J. Phys. Chem. A 119 (2015) 9037–9042. Oxford, 1953.
[25] J.F. Gomes, K. Bergamaski, M.F.S. Pinto, P.B. Miranda, Reaction intermediates of [45] L.N. Konda, I. Czinkota, G. Füleky, G. Morovján, Modeling of single-step and
ethanol electro-oxidation on platinum investigated by SFG spectroscopy, J multistep adsorption isotherms of organic pesticides on soil, J. Agr. Food Chem.
Catal 302 (2013) 67–82. 50 (2002) 7326–7331.
[26] R.M.L. Savedra, M.F.S. Pinto, M. Trsic, E. T., K. Y., Quantum chemical study of [46] D. Do, Adsorption analysis: equilibria and kinetics, Imperial College Press,
electronic and structural properties of retinal and some aromatic analogs, J. London, 1998.
Chem. Phys. 125 (2006) 144901. [47] A.L. Myers, J.M. Prausnitz, Thermodynamics of mixed-gas adsorption, AIChE J.
[27] R. Ditchfield, W.J. Hehre, J.A. Pople, Self-consistent molecular-orbital methods. 11 (1965) 121–127.
IX. An extended gaussian-type basis for molecular-orbital studies of organic [48] C.J. Radke, J.M. Prausnitz, Thermodynamics of multi-solute adsorption from
molecules, J. Chem. Phys. 54 (1971) 724–728. dilute liquid solutions, AIChE J. 18 (1972) 761–768.
[28] E.F. Pettersen, T.D. Goddard, C.C. Huang, G.S. Couch, D.M. Greenblatt, E.C. [49] E. Costa, J.L. Sotelo, G. Calleja, C. Marrón, Adsorption of binary and ternary
Meng, T.E. Ferrin, UCSF Chimera—a visualization system for exploratory hydrocarbon gas mixtures on activated carbon: experimental determination
research and analysis, J. Comput. Chem. 25 (2004) 1605–1612. and theoretical prediction of the ternary equilibrium data, AIChE J. 27 (1981)
[29] B.C.S. Ferreira, F.S. Teodoro, A.B. Mageste, L.F. Gil, R.P. de Freitas, L.V.A. Gurgel, 5–12.
Application of a new carboxylate-functionalized sugarcane bagasse for [50] A.L. Myers, Activity coefficients of mixtures adsorbed on heterogeneous
adsorptive removal of crystal violet from aqueous solution: kinetic, surfaces, AIChE J. 29 (1983) 691–693.
equilibrium and thermodynamic studies, Ind. Crop. Prod. 65C (2015) 521–534. [51] O. Talu, I. Zwiebel, Multicomponent adsorption equilibria of nonideal
[30] Y.S. Ho, G. McKay, Pseudo-second order model for sorption processes, Process mixtures, AIChE J. 32 (1986) 1263–1276.
Biochem. 34 (1999) 451–465. [52] A. Erto, A. Lancia, D. Musmarra, A modelling analysis of PCE/TCE mixture
[31] S.Y. Lagergren, Zur theorie der sogenannten adsorption gelöster stoffe, adsorption based on Ideal Adsorbed Solution Theory, Sep. Purif. Technol. 80
kungliga svenska vetenskapsakademiens, Handlingar 24 (1898) 1–39. (2011) 140–147.
[32] Y. Ho, G. McKay, Kinetic model for lead (II) sorption onto peat, Adsorpt. Sci. [53] A. Erto, A. Lancia, D. Musmarra, A Real Adsorbed Solution Theory model for
Technol. 16 (1998) 243–255. competitive multicomponent liquid adsorption onto granular activated
[33] G.E. Boyd, A.W. Adamson, L.S. Myers, The exchange adsorption of ions from carbon, Micropor. Mesopor. Mat. 154 (2012) 45–50.
aqueous solutions by organic zeolites. II. Kinetics, J. Am. Chem. Soc. 69 (1947) [54] L.V.A. Gurgel, R.P. de Freitas, L.F. Gil, Adsorption of Cu(II), Cd(II), and Pb(II) from
2836–2848. aqueous single metal solutions by sugarcane bagasse and mercerized
[34] W.J. Weber, J.C. Morris, Kinetics of adsorption on carbon from solution, J. San. sugarcane bagasse chemically modified with succinic anhydride, Carbohydr.
Eng. Div. 89 (1963) 31–60. Polym. 74 (2008) 922–929.

Vous aimerez peut-être aussi