You are on page 1of 11

Article

Cite This: Ind. Eng. Chem. Res. 2017, 56, 14078-14088 pubs.acs.org/IECR

Cationization of Cellulose Nanofibers for the Removal of Sulfate Ions


from Aqueous Solutions
Muhammad Muqeet,† Hammad Malik,† Rasool Bux Mahar,*,† Farooq Ahmed,‡ Zeeshan Khatri,‡
and Krista Carlson§

U.S.-Pakistan Center for Advanced Studies in Water and ‡Nanomaterials Research Laboratory, Textile Engineering Department,
Mehran University of Engineering and Technology, Jamshoro 76060, Pakistan
§
Department of Metallurgical Engineering, University of Utah, Salt Lake City, Utah 84112, United States
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: In this study, the adsorption properties of


cationized cellulose nanofibers (c-CNF) were examined for
Downloaded via SOOCHOW UNIV on January 16, 2019 at 03:54:39 (UTC).

the removal of sulfate (SO42−) ions from aqueous solutions


under diverse experimental conditions. Nanofiber mats were
fabricated through electrospinning and cationized with 3-
chloro-2-hydroxypropyl trimethylammonium chloride
(CHTAC). The resultant c-CNF with an ammonium content
of 0.134 mmol/g showed the maximum adsorption capacity of
24.5 mg of SO42− per gram of sorbent using a Langmuir
isotherm model. A pseudo-second-order (PSO) kinetic model
was fitted to the adsorption rate data, showing a higher
adsorption rate of 0.0022 mg g−1 min−1. Scanning electron
microscopy (SEM) micrographs revealed the average fiber
diameter to be 280 ± 10 nm, and a BET surface area of 5.04
m2/g was obtained with a BET surface area and porosity analyzer. Fourier transform infrared (FTIR) spectroscopy confirmed the
conversion of cellulose acetate (CA) to cellulose and its subsequent cationization. Furthermore, the consequences of
cationization were evaluated by zeta potential measurements and thermogravimetric analysis (TGA).

1. INTRODUCTION biodegradable, and inexpensive sorbent is considered to be a


Sulfate contamination in surface water is a major concern in feasible solution for the removal of ions from aqueous
both residential and industrial water supplies. When sulfate- solutions.7
contaminated water is used for residential applications (e.g., Zirconium-loaded sorbents are considered to be an effective
drinking water), ingestion can lead to dehydration, gastro- choice for sulfate removal, but there are issues due to the cost
intestinal irritation, and laxative effects if the concentration of zirconium metal.5 Cellulose has been reported extensively as
becomes greater than the permissible limit (i.e., 250 mg/L).1 a sorbent for aqueous pollutants because of its low cost, as well
Increasing sulfate levels in water provide a strong indication of as its sustainable and ecofriendly nature.8 Past sulfate
the acidification of bodies of water owing to water pollution, adsorption studies have reported different materials such as
which also causes health concerns.2 The use of sulfate-laden coir pitch carbon,9 modified raw straw,5 sugar cane bagasse
water in industrial applications can produce hard scale in heat cellulose,10 surfactant-modified palygorskite,3 carbon residue
exchangers, which ultimately decreases the heat-transfer rate. from biomass gasification,11 cationic cellulose nanofibers from
Additionally, SO42− ions are corrosive in nature, which can lead waste pulp residues,12 and a poly(vinyl chloride) (PVC)−
to material deterioration. The presence of sulfate ions in water zeolite nanoparticle−surfactant anion-exchange membrane.4
is primarily a result of chemical weathering, wide industrial These materials have limitations in terms of extended
application of sulfuric acid, and acidic deposition due to the use adsorption times required for the removal of sulfate ions,
of coal and subsequent oxidation of minerals that contain preparation method, or material costs.
sulfur.3,4 The use of electrospun cationized cellulose nanofibers (c-
Current methods and materials for the removal of sulfate CNF) provides benefits over other absorption methods, as the
ions from water tend to be limited for different reasons or fabrication process is easy, scalable, and cost-effective.13,14 On
costly. For example, biological and ion-exchange treatments are
costly, whereas chemical precipitation methods produce sludge, Received: September 8, 2017
which requires secondary treatment.5 Membrane filtration Revised: October 30, 2017
(mostly reverse osmosis) is highly effective, but it has high Accepted: November 4, 2017
energy consumption.6 In contrast, adsorption with an effective, Published: November 5, 2017

© 2017 American Chemical Society 14078 DOI: 10.1021/acs.iecr.7b03739


Ind. Eng. Chem. Res. 2017, 56, 14078−14088
Industrial & Engineering Chemistry Research Article

Figure 1. Schematic showing the route for the synthesis of cationic cellulose nanofiber sorbent.

the other hand, the remarkable performance of c-CNF within a nm) was received from Archroma Pakistan Ltd. To adjust the
possibly shorter equilibrium time (i.e., 75 min) might become pH for adsorption experiments, 0.1 M HCl or NaOH was used.
meaningful in industrial applications as it would save operating 2.2. Preparation of c-CNF Sorbent. CA solution (17 wt
time compared to other effective adsorbents.15 Additionally, the %) was prepared in 2:1 acetone/DMF and filled in a plastic
fabrication process produces mats, which intrinsically immo- syringe as reported in the literature.23 Briefly, a high-voltage
bilizes the fibers, thereby enabling them to be reused. power supply (HV35P OV, FNM Co. Ltd.) was used under an
Immobilization is a critical challenge to the implementation electric field of 14 kV to overcome surface tension forces with a
of nanostructured materials in practical applications, as any 10-cm tip-to-collector distance. Nanofibers discharged from the
sorbent medium needs to be easily separated from aqueous syringe tip having an internal diameter of 0.6 mm were
solutions after use. An immobilized matrix also enables the collected on a grounded rotating metallic collector (RMC)
potential for reuse, which is important from environmental and wrapped in aluminum foil. The speed of the RMC was
economic perspectives.16,17 controlled at about 12 rpm. The jet was elongated and dried
Several methods have been reported for the fabrication of before it touched the collector, where it formed a ca. 50−70-
polymeric nanofibers, such as drawing,18 template synthesis,19 μm-thick well-connected mat of nanofibers.
self-assembly,20 phase inversion,21 and electrospinning.22 The fabricated web of cellulose acetate nanofibers (CANF)
Among these methods, electrospinning is emerging as a was deacetylated in 0.05 M NaOH aqueous solution for 30 h to
reliable, scalable, and simple technique, because it produces convert them into cellulose nanofibers (CNF). The sample was
smooth and continuous nanofibers with controllable physical air-dried and completely washed with deionized (DI) water to
properties from a variety of polymeric solutions.22,23 remove unfixed −OH− groups.23−26 To prepare c-CNF, the
This work aimed to introduce an efficient, cost-effective, and dried CNF web was again treated with 0.35 M NaOH (first
ecofriendly material for the removal of sulfate ions by the bath) and 1 M CHTAC (second bath) in a two-bath pad-batch
adsorption technique. Kinetic and isotherm modeling of process.24,25 The complete reaction occurred in two steps: (1)
experimental sorption data as a function of time and initial CHTAC was immediately converted into epoxy trimethylam-
concentration was performed to predict the adsorption monium chloride (EPTAC), which (2) subsequently reacted
mechanism and capacity. On the other hand, several with cellulose to form cationic cellulose.25 The two sequential
instrumental techniques were employed to characterize the baths were 100% wet-on-wet. After the second bath, the sample
sorbent, including scanning electron microscopy (SEM), was left in an airtight pot at room temperature for 24 h to allow
Brunauer−Emmett−Teller (BET) analysis, Fourier transform ionic cross-linking to occur, and afterward, it was thoroughly
infrared (FTIR) spectroscopy, thermogravimetric analysis washed with DI water. A graphical depiction of this preparation
(TGA), and zeta potential measurements. process is provided in Figure 1.
2.3. Characterization. Scanning electron microscopy
(SEM) was employed to examine the morphology of the
2. EXPERIMENTAL SECTION
electrospun nanofibers. Samples were sputter-coated (BAL-
2.1. Materials and Reagents. Cellulose acetate (CA, 39.8 TEC AG, Balzers, Liechtenstein) with a 10-nm layer of gold
wt % acetyl content and 30 kDa average molecular weight) was and imaged using a Helios NanoLab scanning electron
obtained from Sigma-Aldrich Japan. N,N-Dimethylformamide microscope at an accelerating voltage of 3 kV and a working
(DMF) and acetone were used as solvents. The cationizing distance of 4.2 mm. The specific surface areas of the CNF and
agent, 3-chloro-2-hydroxypropyl trimethylammonium chloride c-CNF mats were analyzed using a BET surface area and
(CHTAC; average molecular weight 188.1 g/mol, 69 wt % porosity analyzer (Micromeritics ASAP-2020), and the pore
concentration), was obtained from Sigma-Aldrich Japan. C.I. size distribution was determined by nitrogen (N2) adsorption−
Acid Blue 117 dye (molecular weight = 594.5 g/mol, λmax = 580 desorption isotherm. Both samples were degassed individually
14079 DOI: 10.1021/acs.iecr.7b03739
Ind. Eng. Chem. Res. 2017, 56, 14078−14088
Industrial & Engineering Chemistry Research Article

Figure 2. (a−f) SEM micrographs and (b,d,f insets) histograms of (a,b) CANF, (c,d) CNF, and (e,f) c-CNF.

before analysis for 6 h at 90 °C to remove moisture from the S1, Supporting Information). The residual dye concentration
sample tube. was determined by UV/vis/NIR (UV-3600, Shimadzu) spec-
The success of the reaction and chemical modification was troscopy at λmax = 580 nm. Zeta potential measurements of
determined using a Fourier transform infrared (FTIR) CNF and c-CNF were performed at different pH values using a
spectrometer (Nilcolet iS50, Thermo Scientific) in attenu- Zetasizer NanoZS instrument (Malvern, U.K.). For sample
ated-total-reflectance (ATR) mode in the range of wave- analysis, a suspension of ca. 0.1 wt % was prepared in DI water.
numbers from 4000 to 400 cm−1. Another qualitative test was 2.4. Adsorption Studies. A series of preliminary experi-
also employed in this study in which the conversion of CANF ments were performed for isotherm and kinetic modeling to
to CNF was examined through the DMF solubility test as predict the adsorption mechanism and understand the
discussed in the literature.27 DMF is a good solvent for adsorption behavior. In all experiments, 20 mL of aqueous
cellulose acetate, but cellulose cannot be dissolved in this solution was shaken (Lab-line Industries, Inc.) at 200 rpm at ca.
solvent. 23 °C. Adsorption kinetic experiments were carried out by
The thermal stability of CNF and c-CNF was studied to following the batch adsorption method for various contact
determine the changes in chemical and physical properties as a times (15, 30, 45, 60, 75, 90, 105, 120, 135, and 150 min). The
function of increasing temperature at a constant heating rate of sulfate solution was synthesized in DI water with pH
10 °C/min, before and after reaching isothermal conditions adjustments performed through the addition of either 0.1 M
(i.e., 105 °C for 20 min). The data were obtained using an NaOH or HCl.28 Adsorption studies were also carried out as a
SDT-Q600 thermogravimetric analyzer (TA Instruments) function of pH (at values of ca. 5, 6, 7, 8, and 9) and initial
under a nitrogen flow rate of 50 mL/min from 30 to 500 °C. concentration of SO42− ions (at values of 25, 50, 100, 150, 200,
The degree of cationization was determined by an analytical 250, and 300 mg/L).
method in which Acid Blue 117 dye (molecular weight = 594.5 The percentage sulfate removal was determined by analyzing
g/mol), which contains sulfonate groups, was adsorbed using the initial and residual SO42− ion concentrations in water by
CNF and c-CNF at equilibrium conditions (shown in Figure UV/vis/NIR (UV-3600, Shimadzu) spectroscopy to measure
14080 DOI: 10.1021/acs.iecr.7b03739
Ind. Eng. Chem. Res. 2017, 56, 14078−14088
Industrial & Engineering Chemistry Research Article

the absorbance of the aqueous solution (filtrate) at a fixed


wavelength (λ max = 420 nm) and then applying the
turbidimetric method.3 In the turbidimetric method, residual
sulfate ions are precipitated in an acetic acid medium with
barium chloride (BaCl2) to produce barium sulfate (BaSO4)
crystals.29 The light absorbance of the barium sulfate
suspension was measured, and the sulfate concentration was
determined by comparison of the turbidity reading with a stable
calibration curve having R2 = 0.999. On the other hand, the
chloride concentration was critically investigated using the
silver nitrate (AgNO3) titration method to understand the ion-
exchange phenomenon.30

3. RESULTS AND DISCUSSION


3.1. Sorbent Characterization. Physical examination of all
of the fibers is important in the investigation of the effects of Figure 3. FTIR spectra of (a) CANF, (b) CNF, and (c) c-CNF.
the process parameters and external conditions on the
morphology during electrospinning and chemical modification.
Figure 2 shows SEM micrographs of CANF, CNF, and finally functionalization, the c-CNF spectrum (Figure 3c) shows a
prepared c-CNF in which the nanofibers appear to be noticeable new peak at about 1450 cm−1, indicative of CN,
continuous and to exhibit a bead-free morphology. The which is attributed to the presence of quaternary ammonium
diameter of the originally fabricated cellulose acetate nanofibers groups.23
(CANF) ranged from 75 to 650 nm (average diameter = 280 ± The TGA thermograms (Figure 4a,b) provide evidence that
10 nm), and no significant change was observed in the fiber CNF and c-CNF start to degrade at temperatures greater than
appearance or in the diameter of the fibers after deacetylation 200 °C. The first weight loss in both samples was about 3−5%
and cationization. However, slight changes appeared after up to isothermal conditions (i.e., 105 °C for 20 min) owing to
deacetylation in terms of the fiber diameter and surface the release of significant amounts of moisture present in both
roughness. The fiber diameter decreased slightly, and the CNF and c-CNF.34 CNF lost ca. 92% of its weight in the range
micrographs show a rough surface, but, again, the fibers from 220 to 480 °C owing to molecular disintegration into a
retained their surface smoothness after cationization. great number of volatiles and char. The thermogram for CNF
The BET surface areas of CNF and c-CNF are reported in in Figure 4a shows the main degradation peak at ca. 345 °C.
Table 1. A decrease in the surface area occurred after the This degradation peak was mainly attributed to cellulose. The
thermogram in Figure 4b, on the other hand, shows other
Table 1. BET Surface Areas, Pore Volumes, And Zeta degradation peaks at ca. 220 and 280 °C that can be associated
Potentials of the Sorbents at Neutral pH and Ammonium with the structural heterogeneity of c-CNF.12 The thermal
Content in c-CNF degradation analysis indicates that c-CNF can be used at
temperatures up to 220 °C.
zeta
BET surface pore volume potential ammonium The ammonium content (0.134 mmol/g) in c-CNF was
sorbent area (m2/g) (cm3/g) (mV) content (mmol/g) determined by an analytical method (Supporting Information,
CNF 15.20 0.0230 −26.4 − section S1). In addition, surface charge information was
c-CNF 5.04 0.0017 39.1 0.134 determined by zeta potential analysis. In Figure 4c, a shift of
the zeta potential from negative to positive can be seen. A
negative zeta potential corresponds to a negatively charged
cationization of CNF. One potential reason for this decrease is surface and vice versa. CNF presents a negative zeta potential
CHTAC agglomeration owing to interparticle hydrophobic throughout the analysis range, with values ranging from −26.4
interactions,31 which could result in pore blockage and hinder to −10 mV, and its charge values are significantly dependent on
the accessibility of nitrogen during testing.8 Evidence for this the pH of the CNF suspension. The values of the CNF zeta
blockage is also shown in the reduction of the cumulative pore potential over a wide range of pH are attributed to the
volume from 0.0230 cm3/g for CNF to 0.0017 cm3/g for c- negatively charged surface groups (i.e., hydroxyl groups). On
CNF. the other hand, the results revealed a positive zeta potential for
The DMF solubility tests qualitatively confirmed the c-CNF with values ranging from 31.9 to 39.3 mV, as expected
conversion of CANF to CNF, as CNF did not dissolve upon because of the presence of quaternary ammonium functional
contact with DMF, indicating the complete removal of the groups.35 These results agree with the presence of positively
acetyl group. This conversion was also confirmed by the ATR- charged ammonium groups on the surface of c-CNF.
FTIR spectra (Figure 3), which also showed the subsequent 3.2. Comparative Sorption Study. Before selecting a
cationization of CNF to c-CNF. Figure 3a shows three intense sorbent, one must consider different factors such as
peaks matching the stretching vibrations of CO, CCH3, effectiveness, capital costs, design simplicity, flexibility. and
and COC at 1748, 1370, and 1235 cm−1, respectively.13 ease of separation after adsorption. Adsorption data collected
After deacetylation of CANF to CNF (Figure 3b), the carbonyl on c-CNF were compared with previously reported data
absorption peak at 1748 cm−1 disappeared, and the OH (Table. 2) obtained for different sorbent materials such as rice
stretching peak at 3400 cm−1 broadened.32,33 According to straw,5 palygorskite,3 zirconium-loaded sugar cane bagasse
Ahmed et al.,27 this broader OH stretching peak confirms the cellulose,10 carbon residue from biomass gasification,11 coir
complete conversion of cellulose acetate into cellulose. After pitch carbon,9 and c-CNF obtained from pulp residue.12 A mat
14081 DOI: 10.1021/acs.iecr.7b03739
Ind. Eng. Chem. Res. 2017, 56, 14078−14088
Industrial & Engineering Chemistry Research Article

Figure 4. (a,b) TGA thermograms with differential curves of (a) CNF and (b) c-CNF. (c) Zeta potentials of CNF and c-CNF at different pH values.

Table 2. Adsorption Data for the c-CNF Membranea


material adsorption capacity (mg/g) pH equilibrium time (min) ref
c-CNF 24.5 6.8 75 present study
cellulose nanofibers from waste pulp 52.0 (modified) 7.7 ∼720 (overnight) 12
carbon residue from biomass gasification 7.59 (activated carbon) 2 1440 11
19.5 (modified CR) 4
rice straw 11.68 (raw) 6.4 120 5
74.76 (modified)
palygorskite 0.324 (natural) 360 3
0.324 (acid-activated) −
3.240 (surfactant-modified) 4
sugar cane bagasse cellulose 57.5 (zirconium-loaded) 10 5 10
coir pitch carbon 0.06 (no ZnCl2) − 30 9
4.9 (ZnCl2-activated) 4
a
Additional data from other studies included for comparison.

of entangled fibers is advantageous from the standpoint of use chloride and sulfate ions using CNF and c-CNF. At a constant
and reuse because, in practical implementations, it is difficult to sulfate ion concentration (50 mg/L) and time (75 min), it was
separate free-standing nanostructured materials from aqueous found that electrospun CNF showed 18% removal of sulfate
solutions.43 Although several authors have reported higher ions and 11% removal of chloride ions. Similarly, c-CNF
adsorption capacities for other materials and/or preparation showed 75% and 32% removals of sulfate and chloride ions,
methods compared to the electrospun nanofibers developed in respectively, under the same equilibrium conditions (see Figure
this work, there are several drawbacks to those materials that 5a). This difference in adsorption was due to the cationic
could limit their use in commercial settings, such as extended charge strength of the ammonium groups on c-CNF, and the
adsorption times and high material costs. prepared c-CNF membrane was found to be more selective
3.3. Binding Mechanism of Sulfate Ions. The adsorption toward divalent sulfate ions.12
behavior and binding mechanism of sulfate ions on the sorbent Bearing in mind the problem of disposing of used adsorbent
used in this work were explored by comparing the adsorption of and minimizing environmental pollution due to waste disposal,
14082 DOI: 10.1021/acs.iecr.7b03739
Ind. Eng. Chem. Res. 2017, 56, 14078−14088
Industrial & Engineering Chemistry Research Article

Figure 5. (a) Comparative adsorption removals of sulfate and chloride. (b) Schematic map showing the electrostatic relationship between negatively
charged sulfate ions and positively charged ammonium groups grafted on the surface of cationic cellulose nanofibers.

recycling of saturated c-CNF after the first trial of the PFO model
adsorption test was attempted. For this purpose, saturated c- k1
CNF was regenerated using 0.35 M NaOH (for sulfate log(qe − qt ) = log qe − t
2.303 (1)
desorption), and ca.1−2 mg of fresh c-CNF was added to
make up for the weight lost during adsorption and the fixed PSO model
weight of nanofibers (i.e., 10 mg), as well as stimulating the t 1 t
used materials with new active sites. The experimental = +
qt k 2qe 2 qe (2)
observations showed that the regenerated cellulose nanofibers
were capable of removing 55% of the sulfate ions present in the where qt and qe represent the amounts adsorbed at time t and at
aqueous solution under the optimized conditions (pH 6.8; equilibrium, respectively, and k1 and k2 represent the rate
initial concentration, 50 mg/L; contact time, 75 min). coefficients for the PFO and PSO kinetic models,
3.4. Adsorption Kinetics. It was perceived from the kinetic respectively.44
data that the adsorption was rapid in the first 15 min and Solid−liquid adsorption generally involves film diffusion and
afterward progressed at the comparatively lower rate because of intraparticle diffusion (IPD).36 Neither the PFO nor PSO
kinetic model describes this diffusion phenomenon. Therefore,
the sufficient numbers of vacant active sites available at the
the IPD model (Figure 6d) was constructed to define the
initial times on the surface of sorbent.44 At the 75-min process of diffusion. According to this model, a plot of qt versus
equilibration time, the surface of the sorbent became saturated, t1/2 must be linear and pass through the origin for IPD to be the
and subsequently, the removal efficiency became almost rate-limiting step. In the model developed by Poots, Morris,
constant. This equilibrium time was selected for further and McKay, the initial rate of IPD is calculated by linearization
adsorption studies. of the equation
Pseudo-first-order (PFO; Figure 6b) and pseudo-second-
qt = K it 0.5 + C (3)
order (PSO; Figure 6c) kinetic models describe the adsorption
rate and the factors that influence the rate in the attainment of where Ki represents the IPD rate constant and C is the
equilibrium in a reasonable amount of time.36,37 The equations intercept. The value of C (intercept) gives an idea of the
for these models can be expressed as follows thickness of the boundary layer; that is, the larger the value of
14083 DOI: 10.1021/acs.iecr.7b03739
Ind. Eng. Chem. Res. 2017, 56, 14078−14088
Industrial & Engineering Chemistry Research Article

Figure 6. (a) Plot of sulfate uptake kinetics by c-CNF (pH ∼7; mass of sorbent, 10 mg; initial concentration, 50 mg/L; shaking time range, 15−150
min). (b−d) Fits to (b) pseudo-first-order (PFO), (c) pseudo-second-order (PSO), and (d) intraparticle diffusion (IPD) kinetic models.

Table 3. Kinetic Parameters of Pseudo-First-Order, Pseudo-Second-Order, and Intraparticle Diffusion Rate Equations
pseudo-first-order pseudo-second-order intraparticle diffusion
parameter value parameter value parameter value
−1 −1 −1 −1 −1/2
K1 (min ) 0.0285 K2 (mg g min ) 0.0022 Ki (mg g min ) 1.68
qe (mg/g) 22.98 qe (mg/g) 62.11 C 40.64
R2 0.673 R2 0.995 R2 0.76

C, the greater the boundary layer effect.38 The value of the IPD IPD starts to become slower because of the low sulfate ion
rate constant (see Table 3) clearly shows that the boundary concentration in the aqueous solution.40
layer has a significant effect on the diffusion mechanism of 3.5. Adsorption Isotherm. As the concentration of sulfate
sulfate ions uptake by c-CNF. ions present in natural water varies depending on location, it is
The results reported in Table 3 provide the values of imperative to investigate the adsorption capacity of the sorbent
constants obtained by kinetic modeling. These results indicate at different concentrations. Figure 7 illustrates the adsorption of
that the adsorption of sulfate ions follows PSO kinetics, sulfate ions as a function of initial concentration. Figure 7a
suggesting the co-occurrence of both physisorption and shows the reduction in adsorption efficiency of c-CNF with
increasing initial concentration of sulfate ions. This reduction
chemisorption.39,40 Figure 6d shows that a plot of qt versus
can be attributed to the finite number of active sites available on
t1/2 is not linear over the whole range of time. Rather, the data c-CNF that were occupied by adsorbed sulfate ions,
points exhibit a trilinearity of three successive adsorption stages subsequently leading to a decrease in sulfate adsorption.
of mass transport with a decreasing rate. This type of trilinearity The experimentally derived batch adsorption data were also
can be attributed to (1) adsorption on the outer surface correlated with the two most famous adsorption isotherm
associated with boundary layer diffusion, (2) IPD states where models, namely, the Langmuir and Freundlich models, to
this step is highly involved in the rate control of this determine the adsorption mechanism.7 The equations for the
mechanism, and (3) the ultimate equilibrium stage where Langmuir and Freundlich models are given by
14084 DOI: 10.1021/acs.iecr.7b03739
Ind. Eng. Chem. Res. 2017, 56, 14078−14088
Industrial & Engineering Chemistry Research Article

Figure 7. (a) Effects of initial concentration on the adsorption of sulfate ions (pH ∼7; mass of sorbent, 10 mg; shaking time, 75 min; concentration
range, 25−300 mg/L). (b−d) Fits to (b) nonlinear adsorption isotherms, (c) the Langmuir isotherm, and (d) the Freundlich isotherm.

Langmuir model the sorbent (c-CNF) and the concentration of residual solute in
Ce C solution after adsorption.
1
= e + The essential characteristic term of the Langmuir equation
qe qm qmKL (4) can be expressed by RL, which is given by
Freundlich model RL = 1/(1 + KLC0) (6)
1
ln qe = ln Ce + ln KF where RL designates the type of isotherm, which can be linear
n (5)
(RL = 1), irreversible (RL = 0), favorable (0 < RL < 1), or
where qm is the maximum adsorption capacity; Ce is the unfavorable (R L > 1), and C 0 represents the initial
equilibrium solution-phase concentration; KL is the Langmuir concentration of sulfate ions.
constant, which is associated with the adsorption free energy Table 4 shows that the Langmuir model fit the adsorption
and specifies the sorbent−sorbate affinity; KF is the Freundlich isotherm parameters better than the Freundlich model.
constant, which gives the adsorption capacity at unit Moreover, the maximum experimental adsorption value (i.e.,
concentration; and the value of 1/n in the Freundlich
isotherm describes the comparative dissemination of active Table 4. Parameters of the Langmuir and Freundlich
sites: the smaller the value of 1/n, the greater the availability of Adsorption Isotherms for Sulfate (SO42−) Adsorption
heterogeneous active sites and the greater the likelihood that
the adsorption mechanism is physical in nature.15 The Langmuir isotherm Freundlich isotherm
Langmuir adsorption isotherm model assumes that adsorption
constant value constant value
of a monolayer occurs with no interactions between adjacent
adsorbed molecules on identical surface sites and that the same qexp (mg/g) 27.32 1/n 0.229
mechanism is involved for all adsorption.9,45 This contrasts with qm,cal (mg/g) 24.5 Kf (mg/L) 2.076
the Freundlich model, which is an empirical equation and the KL (L/mg) 0.432 R2 0.742
linear form of this isotherm model. This is an empirical RL 0.0076
relationship between the solute (sulfate ion) concentration on R2 0.990

14085 DOI: 10.1021/acs.iecr.7b03739


Ind. Eng. Chem. Res. 2017, 56, 14078−14088
Industrial & Engineering Chemistry Research Article

qexp = 27.32 mg/g) is very close to the calculated value of qm, 4. CONCLUSIONS
the maximum theoretical adsorption value of sulfate ion uptake In this study, nanostructured cationic cellulose nanofibers (c-
for c-CNF. The current results are evidence for the formation CNF) with an average diameter of 280 ± 10 nm and an
of a monomolecular layer of sulfate ions on the surface of c- ammonium content of 0.134 mmol/g were successfully
CNF, which also confirms that chemisorption is responsible for prepared. The chemical alteration reversed the surface charge
the uptake of sulfate ions.46,47 of CNF to positive values and revealed the other thermal
3.6. pH Effect on Sulfate Adsorption. The pH value of degradation peaks of CHTAC-modified cellulose in TGA
an aqueous solution has a strong effect on the adsorption experiments. Owing to the positive charges, c-CNF showed a
efficiency of a sorbent. Figure 8 displays the sulfate removal promising capability to adsorb negatively charged sulfate ions
through electrostatic interactions and complexation. It was
concluded that the Langmuir model fit the data well, presenting
a high value of R2 = 0.99, and that the experimental adsorption
data followed a pseudo-second-order kinetic model, indicating
the co-occurrence of both physisorption and chemisorption.
The maximum adsorption capacity of c-CNF was determined
to be 24.5 mg of sulfate ions per gram of sorbent.

■ ASSOCIATED CONTENT
* Supporting Information
S
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acs.iecr.7b03739.
Description of the method used to determine the degree
of cationization, UV−vis spectra of Acid Blue 117 dye,
and TGA curve representing the percentage weight loss
(PDF)

Figure 8. Plot of percentage sulfate removals at different pH values


■ AUTHOR INFORMATION
Corresponding Author
(initial concentration, 50 mg/L; mass of sorbent, 10 mg; shaking time, *E-mail: rbmahar.uspcasw@faculty.muet.edu.pk. Tel.: +92-22-
75 min). 2772250 ext. 6508.
ORCID
percentages obtained using c-CNF. The sulfate adsorption Rasool Bux Mahar: 0000-0002-1966-0397
behavior in near-neutral pH solutions is based on the Zeeshan Khatri: 0000-0001-8779-3805
interaction of the positively charged ammonium groups with Notes
the negatively charged sulfate ions present in the aqueous The authors declare no competing financial interest.
solution.41 A maximum removal of ca. 73% was achieved at pH
7 because of the number of factors employed in the present
adsorption such as hydrogen bonding, electrostatic interactions,
■ ACKNOWLEDGMENTS
The authors are grateful to the U.S.-Pakistan Center for
and complexation.43 Another potential reason for this behavior Advanced Studies in Water, Mehran University of Engineering
can also be provided by the Hofmeister effect if the solution and Technology, Jamshoro, Sindh, Pakistan. for providing
contains nonpolar molecules.48 However, in this study, financial support. We also acknowledge the Roger and Dawn
adsorption experiments were conducted in synthetically Crus Center for Renewable Energy at the University of Utah,
prepared sulfate solutions. Salt Lake City, UT, for technical assistance and sample
At lower pH values, the SO42− ions present in aqueous characterization.


solution are protonated to form hydrogen sulfate or bisulfate
ions (HSO4−). Consequently, there will be a low force of REFERENCES
attraction between the positively charged surface of the sorbent (1) Guidelines for Drinking Water Quality, 4th ed.; World Health
and the protonated sulfate ions, resulting in a reduction of the Organization: Geneva, Switzerland, 2011.
percentage adsorption.43,46 On the other hand, as the pH is (2) Nezamzadeh-Ejhieh, A. N.; Esmaeilian, A. Application of
increased, the concentration of H+ decreases, and at the same surfactant modified zeolite carbon paste electrode (SMZ-CPE)
time, the concentration of OH− increases. Thus, OH− and towards potentiometric determination of sulfate. Microporous Meso-
SO42− coexist with the same electrostatic charges and compete porous Mater. 2012, 147, 302.
for the same adsorption sites available on c-CNF. This is the (3) Dong, R.; Liu, Y.; Wang, X.; Huang, J. Adsorption of Sulfate Ions
reason for the ddecrease in the adsorption efficiency with from Aqueous Solution by Surfactant-Modified Palygorskite. J. Chem.
increasing pH.9,42 In addition, the presence of Na+ ions from Eng. Data 2011, 56, 3890.
(4) Danesh Khorasgani, M.; Nezamzadeh-Ejhieh, A. N. PVC-zeolite
the NaOH titrant could have produced a transmission effect, nanoparticle-surfactant anion exchanger membrane: preparation,
which is responsible for a reduction in removal efficiency.35 Still characterization, and its application in development of ion-selective
for lower and higher pH values, the removal percentages of electrode for detection of sulfate. J. Solid State Electrochem. 2016, 20,
sulfate ions are greater than 50%. This indicates the influence of 2827.
physical adsorption due to hydrogen bonding and the interpore (5) Cao, W.; Dang, Z.; Zhou, X.-Q.; Yi, X.-Y.; Wu, P.-X.; Zhu, N.-W.;
structure of the nanofibers. Lu, G.-N. Removal of sulphate from aqueous solution using modified

14086 DOI: 10.1021/acs.iecr.7b03739


Ind. Eng. Chem. Res. 2017, 56, 14078−14088
Industrial & Engineering Chemistry Research Article

rice straw: Preparation, characterization and adsorption performance. (25) Hashem, M.; Hauser, P.; Smith, B. Reaction Efficiency for
Carbohydr. Polym. 2011, 85, 571. Cellulose Cationization Using 3-Chloro-2- Hydroxypropyl Trimethyl
(6) Mbakop, S.; Mthombeni, N. H.; Leswifi, T. Y.; Onyango, M. S. Ammonium Chloride. Text. Res. J. 2003, 73, 1017.
Evaluation of Chitosan-Bentonite Composite Performance towards (26) Gopiraman, M.; Bang, H.; Yuan, G.; Yin, C.; Song, K.-H.; Lee, J.
Remediation of Sulphate Containing Effluent. Proc. Sustainable Res. S.; Chung, I. M.; Karvembue, R.; Kim, I. S. Noble metal/functionalized
Innovation Conf. 2016, 65−70. cellulose nanofiber composites for catalytic applications. Carbohydr.
(7) Chen, X. Modeling of Experimental Adsorption Isotherm Data. Polym. 2015, 132, 554.
Information 2015, 6, 14. (27) Ahmed, F.; Ayoub Arbab, A.; Jatoi, A. W.; Khatri, M.; Memon,
(8) Sehaqui, H.; Gálvez, M. E.; Becatinni, V.; Ng, Y.; Steinfeld, A.; N.; Khatri, Z.; Kim, I. Ultrasonic-assisted deacetylation of cellulose
Zimmermann, T.; Tingaut, P. Fast and Reversible Direct CO2 Capture acetate nanofibers: A rapid method to produce cellulose nanofibers.
from Air onto All-Polymer Nanofibrillated Cellulose-Polyethylenimine Ultrason. Sonochem. 2017, 36, 319.
Foams. Environ. Sci. Technol. 2015, 49, 3167. (28) Silva, R.; Cadorin, L.; Rubio, J. Sulphate ions removal from an
(9) Namasivayam, C.; Sangeetha, D. Application of coconut coir pith aqueous solution: I. Co-precipitation with hydrolysed aluminum-
for the removal of sulfate and other anions from water. Desalination bearing salts. Miner. Eng. 2010, 23, 1220.
2008, 219, 1. (29) Kortazar, L.; Saez, J.; Agirre, J.; Izaguirre, J. K.; Fernandez, L. A.
(10) Mulinari, D. R.; da Silva, M. L. Adsorption of sulphate ions by Analytical Methods Application of multivariate analysis to the
modification of sugarcane bagasse cellulose. Carbohydr. Polym. 2008, turbidimetric determination of sulphate in seawater. Anal. Methods
74, 617. 2014, 6, 3510.
(11) Runtti, H.; Tuomikoski, S.; Kangas, T.; Kuokkanen, T.; Rämö, (30) Krishna Kumar, S.; Bharani, R.; Magesh, N. S.; Godson, P. S.;
J.; Lassi, U. Sulphate Removal from Water by Carbon Residue from Chandrasekar, N. Hydrogeochemistry and groundwater quality
Biomass Gasification:Effect of Chemical Modification Methods on appraisal of part of south Chennai coastal aquifers, Tamil Nadu,
Sulphate Removal Efficiency. BioResources 2016, 11, 3136. India using WQI and fuzzy logic method. Appl. Water Sci. 2014, 4, 341.
(12) Sehaqui, H.; Mautner, A.; Perez de Larraya, U.; Pfenninger, N.; (31) Bhatt, A. S.; Sakaria, P. L.; Vasudevan, M.; Pawar, R.; Sudheesh,
Tingaut, P.; Zimmermann, T. Cationic cellulose nanofibers from waste N.; Bajaj, H.; Mody, H. M. Adsorption of an anionic dye from aqueous
pulp residues and their nitrate, fluoride, sulphate and phosphate medium by organoclays: equilibrium modeling, kinetic and thermody-
adsorption properties. Carbohydr. Polym. 2016, 135, 334. namic exploration. RSC Adv. 2012, 2, 8663.
(13) Khatri, Z.; Arain, R. A.; Jatoi, A. W.; Mayakrishnan, G.; Wei, K.; (32) Chen, X.; Shen, B.; Sun, H.; Zhan, G.; Huo, Z. Adsorption and
Kim, I.-S. Dyeing and characterization of cellulose nanofibers to Its Mechanism of CS2 on Ion-Exchanged Zeolites Y. Ind. Eng. Chem.
improve color yields by dual padding method. Cellulose 2013, 20, Res. 2017, 56, 6499.
1469. (33) Tian, X.; Yan, D.; Lu, Q.; Jiang, X. Cationic surface modification
(14) Mahapatra, A.; Mishra, B. G.; Hota, G. Studies on Electrospun of nanocrystalline cellulose as reinforcements for preparation of the
Alumina Nano fi bers for the Removal of Chromium (VI) and chitosan-based nanocomposite films. Cellulose 2017, 24, 163.
Fluoride Toxic Ions from an Aqueous System. Ind. Eng. Chem. Res. (34) Zhou, Y.; Hu, X.; Zhang, M.; Zhuo, X.; Niu, J. Preparation and
2013, 52, 1554. Characterization of Modified Cellulose for Adsorption of Cd (II), Hg
(15) Qureshi, U. A.; Khatri, Z.; Ahmed, F.; Khatri, M.; Kim, I. (II), and Acid Fuchsin from Aqueous Solutions. Ind. Eng. Chem. Res.
Electrospun Zein Nano fi ber as a Green and Recyclable Adsorbent for 2013, 52, 876.
the Removal of Reactive Black 5 from the Aqueous Phase. ACS (35) Sehaqui, H.; Perez de Larraya, U.; Tingaut, P.; Zimmermann, T.
Sustainable Chem. Eng. 2017, 5, 4340. Humic acid adsorption onto cationic cellulose nanofibers for
(16) Shahbeig, H.; Bagheri, N.; Ghorbanian, S. A.; Hallajisani, A.; bioinspired removal of copper(II) and a positively charged dye. Soft
Poorkarimi, S. A new adsorption isotherm model of aqueous solutions Matter 2015, 11, 5294.
on granular activated carbon. World J. Model. Simul. 2013, 9, 243. (36) Haider, S.; Binagag, F. F.; Haider, A.; Mahmood, A.; Shah, N.;
(17) Sehaqui, H.; Michen, B.; Marty, E.; Schaufelberger, L.; Al-Masry, W. A.; Khan, S.nU.-D.; Ramay, S. M. Adsorption kinetic and
Zimmermann, T. Functional Cellulose Nano fi ber Filters with isotherm of methylene blue, safranin T and rhodamine B onto
Enhanced Flux for the Removal of Humic Acid by Adsorption. ACS electrospun ethylenediamine-grafted polyacrylonitrile nanofibers mem-
Sustainable Chem. Eng. 2016, 4, 4582. brane. Desalin. Water Treat. 2015, 55, 1609.
(18) Suzuki, A.; Shimizu, R. Biodegradable Poly (glycolic acid) (37) Dixit, A.; Mishra, P. K.; Alam, M. S. Titania Nanofibers: A
Nanofiber Prepared by CO2 Laser Supersonic Drawing. J. Appl. Polym. Potential Adsorbent for Mercury and Lead Uptake. Int. J. Chem. Eng.
Sci. 2011, 121, 3078. Appl. 2017, 8, 75.
(19) Wang, S.; Zhang, Y.; Wang, W.; Li, G.; Ma, X.; Li, X.; Zhang, Z.; (38) Babitha, S.; Korrapati, P. S. TiO2 immobilized zein micro-
Qian, Y. Template-assisted synthesis of porous molybdenum dioxide spheres: a biocompatible adsorbent for effective dye decolourisation.
nanofibers and nanospheres by redox etching method. J. Cryst. Growth RSC Adv. 2015, 5, 26475.
2006, 290, 96. (39) Robati, D. Pseudo-second-order kinetic equations for modeling
(20) Subbiah, T.; Bhat, G. S.; Tock, R. W.; Parameswaran, S.; adsorption systems for removal of lead ions using multi-walled carbon
Ramkumar, S. S. J. Appl. Polym. Sci. 2005, 96, 557. nanotube. J. Nanostructure in Chemistry 2013, 3, 55.
(21) Cai, Y.; Hou, X.; Wang, W.; Liu, M.; Zhang, J.; Qiao, H.; Huang, (40) García, E. R.; Medina, R. L.; Lozano, M.; Hernández Perez, I.;
F.; Wei, Q. Effects of SiO2 nanoparticles on structure and property of Valero, M.; Franco, A. Adsorption of Azo-Dye Orange II from
form-stable phase change materials made of cellulose acetate phase Aqueous Solutions Using a Metal-Organic Framework Material: Iron-
inversion membrane absorbed with capric-myristic-stearic acid ternary Benzenetricarboxylate. Materials 2014, 7, 8037.
eutectic mixture. Thermochim. Acta 2017, 653, 49. (41) Liu, P.; Sehaqui, H.; Tingaut, P.; Wichser, A.; Oksman, K.;
(22) Ahmed, F. E.; Lalia, B. S.; Hashaikeh, R. A review on Mathew, A. P. Cellulose and chitin nanomaterials for capturing silver
electrospinning for membrane fabrication: Challenges and applica- ions (Ag + ) from water via surface adsorption. Cellulose 2014, 21, 449.
tions. Desalination 2015, 356, 15. (42) Xu, D.; Zhu, K.; Zheng, X.; Xiao, R. Poly (ethylene-co-vinyl
(23) Khatri, Z.; Mayakrishnan, G.; Hirata, Y.; Wei, K.; Kim, I. alcohol) Functional Nanofiber Membranes for the Removal of Cr (VI)
Cationic-cellulose nanofibers: Preparation and dyeability with anionic from Water. Ind. Eng. Chem. Res. 2015, 54, 6836.
reactive dyes for apparel application. Carbohydr. Polym. 2013, 91, 434. (43) Patel, S.; Hota, G. Iron oxide nanoparticle-immobilized PAN
(24) Pei, A.; Butchosa, N.; Berglund, L. A.; Zhou, Q. Surface nanofibers: synthesis and adsorption studies. RSC Adv. 2016, 6, 15402.
quaternized cellulose nanofibrils with high water absorbency and (44) Anari Anaraki, M.; Nezamzadeh-Ejhieh, A. N. Modification of
adsorption capacity for anionic dyes. Soft Matter 2013, 9, 2047. an Iranian clinoptilolite nano particles by hexadecyl trimethyl

14087 DOI: 10.1021/acs.iecr.7b03739


Ind. Eng. Chem. Res. 2017, 56, 14078−14088
Industrial & Engineering Chemistry Research Article

ammonium cationic surfactant and dithizone for removal of Pb (II)


from aqueous solution. J. Colloid Interface Sci. 2015, 440, 272.
(45) Chaleshtori, M. H.; Nezamzadeh-Ejhieh, A. N. Clinoptilolite
nano-particles modified with aspartic acid for removal of Cu(II) from
aqueous solutions: isotherms and kinetic aspects. New J. Chem. 2015,
39, 9396.
(46) Shirzadi, H.; Nezamzadeh-Ejhieh, A. N. An efficient modified
zeolite for simultaneous removal of Pb (II) and Hg(II) from aqueous
solution. J. Mol. Liq. 2017, 230, 221.
(47) Naghash, A.; Nezamzadeh-Ejhieh, A. N. Comparison of the
efficiency of modified clinoptilolite with HDTMA and HDP
surfactants for the removal of phosphate in aqueous solutions. J. Ind.
Eng. Chem. 2015, 31, 185.
(48) Hofmeister, F. To the doctrine of the action of the salts.
Naunyn-Schmiedeberg's Arch. Pharmacol. 1888, 24, 247.

14088 DOI: 10.1021/acs.iecr.7b03739


Ind. Eng. Chem. Res. 2017, 56, 14078−14088

You might also like