You are on page 1of 12

Clay Minerals (2019), 54, 197–207

doi:10.1180/clm.2019.27

Article

Evaluation of the efficacy of halloysite nanotubes in the removal of


acidic and basic dyes from aqueous solution
Tholiso Ngulube1*, Jabulani Ray Gumbo2, Vhahangwele Masindi3,4 and Arjun Maity5,6
1
Department of Ecology and Resources Management, School of Environmental Sciences, University of Venda, Private Bag X5050, Thohoyandou, 0950, Limpopo,
South Africa; 2Department of Hydrology and Water Resources, School of Environmental Sciences, University of Venda, Private Bag X5050, Thohoyandou, 0950,
Limpopo, South Africa; 3Council for Scientific and Industrial Research (CSIR), Built Environment, Hydraulic Infrastructure Engineering, PO Box 395, Pretoria, 0001,
South Africa; 4Department of Environmental Sciences, School of Agriculture and Environmental Sciences, University of South Africa (UNISA), PO Box 392, Florida,
1710, South Africa; 5Department of Applied Chemistry, University of Johannesburg, Johannesburg, South Africa and 6DST/CSIR National Centre for Nanostructured
Materials, Council for Scientific and Industrial Research (CSIR), Pretoria, South Africa

Abstract
The present work describes the removal of Direct Red 81, Methyl Orange, Methylene Blue and Crystal Violet from aqueous solution
using halloysite nanotubes. The clay mineral was physicochemically characterized using various methods. The influences of pH, inter-
action time, initial dye concentration, adsorbent amount and temperature on adsorption were monitored and interpreted. Although pre-
vious work has shown that acidic pH conditions favour the adsorption of pollutants from aqueous systems by clay materials, in this study
maximum removal was possible over a wide range of pH conditions (pH ≥2–12). Adsorption was very rapid, and equilibrium was
attained within 30 min. For all four dyes studied, chemical reaction seemed significant in the rate-controlling step, and the pseudo-
second-order chemical reaction kinetics provided the best correlation of the experimental data. Thermodynamically, the process was
spontaneous, with Gibbs energy decreasing with increasing temperature. Halloysite would be suitable for removing dyes from aqueous
solution. This was further tested by using the halloysite nanotubes for the removal of complex dyes from printing and ink industry
effluents.

Keywords: halloysite nanotubes, acidic and basic dyes, adsorption, kinetics, isotherms, thermodynamics
(Received 31 August 2018; revised 17 March 2019; Associate Editor: C.-H. Zhou)

Coloured, toxic industrial waste is increasingly becoming a social decontamination, naturally available clays have been applied in
and environmental problem. The contaminants may spread most developing countries (Adebowale et al., 2014; Gamoudi &
through living organisms via aquatic pathways (Ngulube et al., Srasra, 2018; Kausar et al., 2018). It is extremely important to
2017), necessitating the development of numerous techniques to choose ta echnique for water decontamination that is cheap and
eliminate dyes from wastewater (Yang et al., 2018). Adsorption requires minimal workload. Using clay minerals and soils fits
is among the methods used for water treatment because of its well with the above-mentioned requirements.
low start-up costs, straightforward design and ease of operation Halloysite clay is an inexpensive adsorbent applied for the deco-
and implementation, even in small plants (Mu & Wang, 2016). lourization of dyed wastewater. Halloysite (Al2Si2O5(OH)4·nH2O,
Because of stringent laws concerning the removal of dyes from where n is 0 in kaolinite and up to 4 in halloysite) is a dioctahedral
wastewaters before their discharge into the aquatic environment 1:1 clay mineral belonging to the kaolin group (Berthier, 1826). It
and the cost of conventional adsorbents, development of new, has a similar structure to kaolinite, but it is rolled in tubes with dia-
low-cost adsorbents with improved adsorption properties has meters of ∼50 nm and lengths of ∼1000 nm (Słomkiewicz et al.,
become commonplace (Yagub et al., 2014). 2015). Halloysite clay displays greater adsorption capacity for
Various abundant, naturally occurring materials have been both acidic and basic dyes because of the presence of a negatively
examined as possible adsorbents for dye treatment (Santos charged SiO2 outermost surface and positively charged Al2O3
et al., 2016; Sarma et al., 2016). There is increasing focus on nat- inner surface; thus, it has effective bivalent adsorbing capacity
ural adsorbents, and preferably the development of adsorbents (Zhao et al., 2013).
demonstrating both low costs and high adsorption capacities for Most previous studies have focused on an individual dye using
removing inorganic contaminants from polluted waters. synthetic solutions. The objectives of the present study were to
Although a variety of sorbents have been used for water evaluate the adsorption capacity of halloysite clay for two anionic
and two cationic dyes, Direct Red 81 (DR81), Methyl Orange
(MO), Methylene Blue (MB) and Crystal Violet (CV), which
*E-mail: tholisongulube@gmail.com
are used in the textile industry, and to evaluate the influences
Cite this article: Ngulube T, Gumbo JR, Masindi V, Maity A (2019). Evaluation of the
efficacy of halloysite nanotubes in the removal of acidic and basic dyes from aqueous of several parameters on the adsorption capacity of the clay.
solution. Clay Minerals 54, 197–207. https://doi.org/10.1180/clm.2019.27 The adsorption kinetics and isotherms were studied in order to
Copyright © Mineralogical Society of Great Britain and Ireland 2019
198 Tholiso Ngulube et al.

determine the adsorption capacity of the clay. This provided an analysis. The FTIR spectra were recorded using a FTIR spectrometer
opportunity to perform a comparative analysis on how the halloy- (Bruker, Alpha, USA) with a scan range of 450–4500 cm–1.
site material adsorbs variously charged dyes under the same experi- Transmission electron microscopy (TEM) images were acquired
mental conditions. There have been few studies that test adsorbent using a JEM-2100F Field Emission microscope (JEOL, Japan).
materials on real, coloured industrial effluents. In the present study, The point of zero charge (PZC) was determined using the solid
halloysite was tested on real industrial wastewater in order to ascer- addition method described by Izuagie et al. (2016). The Brunauer–
tain the feasibility of the adsorbent under real conditions. Emmett–Teller (BET) specific surface area was determined by N2
adsorption (Micromeritics Tristar II, Norcross, USA).
Materials and methods
Adsorbent regeneration
Materials
The regeneration potential of halloysite was evaluated with 0.01 M
The halloysite nanoclay–kaolin clay was purchased from Sigma NaOH. An adsorption experiment was performed using 2 g of clay,
Aldrich (South Africa). MO, MB, CV, HCl, KCl and NaOH pellets 50 mL of solution and a 10 mg/L concentration of each dye. The
were purchased from Rochelle Chemicals (South Africa). DR81 was amount of dye adsorbed was noted and the adsorbent was dried
purchased from Sigma Aldrich (South Africa). Real industrial effluent in the oven for 12 h at 105°C. The adsorbent was then dispersed
was obtained from an operator in the printing and ink industry. in 50 mL of 0.01 M NaOH solution and the dispersion was centri-
fuged at 50,000 rpm for 15 min. The adsorbent was washed by 50
Batch removal studies mL MilliQ water, dried in an oven for 12 h at 105°C and cooled
in a desiccator. The dried adsorbent was used for another adsorp-
Deionized water (ELGA, Micra Veolia Water Solution and tion experiment following the procedure described above. The
Technologies, UK) was used to prepare dye solutions by dissolv- same procedure using the same adsorbent was repeated three times.
ing appropriate amounts, accurately weighed on an electronic bal-
ance (RADWAG electronic balance, Wagi Elektroniczen, Poland), Results and discussion
of dry powder dye to prepare a stock solution (1000 mg/L).
Experimental solutions were obtained by serial dilutions to obtain Characterization
the working solutions at desired concentrations. The final concen- X-ray diffraction results
trations of the dyes were estimated for each sample absorbance at The XRD traces of the original halloysite and the halloysite after
the wavelength corresponding to the maximum absorption peak adsorption of DR81, MO, MB and CV dyes are shown in Fig. 1.
(λmax = 510, 505, 664 and 590 nm for DR81, MO, MB and CV, There is no significant difference in the traces after adsorption
respectively) using a Vis spectrophotometer (Thermoscientific of the dyes. The diffractograms showed sharp peaks at 13.95,
Orion Aqua Matte 7000, China) using a constructed calibration 23.23 and 31.02°2θ. The results agree with previous reports
graph from prepared standard solutions. Batch adsorption experi- on halloysite nanotubes (Jiang et al., 2015). The peaks identified
ments were performed in 250 mL glass Erlenmeyer flasks contain- are characteristic of quartz (SiO2), kaolinite (Al2Si2O5OH4) and
ing 50 mL of dye solution of desired concentrations at various pH halloysite (Al2Si2O5OH4) (Zhang et al., 2012). The similarity of
values (2, 4, 6, 8, 10 and 12), initial dye concentrations (1, 5, 10, raw halloysite and dye-adsorbed halloysite diffractograms indi-
20, 30 and 40 mg/L), temperatures (25°C, 35°C, 45°C and 55°C), cates that the dye-adsorption process did not alter the crystallog-
contact times (10, 15, 20, 30, 60 and 90 min) and adsorbent raphy or major mineral phases of the sample.
dosages (0.1, 0.2, 0.4, 0.8, 1.0 and 2 g). The solution pH was
adjusted using either 0.1 M HCl or 0.1 M NaOH solution. The FTIR results
solutions were shaken on a reciprocating table shaker (Labotec, The FTIR spectra of raw halloysite and the halloysite after adsorp-
Model 207, South Africa) at 250 rpm. After the equilibration tion of DR81, MO, MB and CV dyes are shown in Fig. 2. The
time, the solutions were allowed to settle for 30 min and the spectra display bands in the OH-stretching region (3700 cm–1)
absorbance of the supernatant solution was recorded. All of the (Luo et al., 2010) and bands associated with stretching and bend-
experiments besides those of temperature variation were carried ing vibrations of the aluminosilicate framework (∼1200 cm–1)
out at room temperature (25 ± 3°C). In order to verify the repro- typical of kaolin minerals (Prokop et al., 2015). The bands at
ducibility of the results, the experiments were carried out in trip- 3696 and 3623 cm–1 are due to the stretching vibrations of
licate and the mean values were reported. hydroxyl groups at the surfaces of the nanotubes (Xie et al.,
2011). Fourier-transform infrared analysis is an important tool
for identifying the characteristic functional groups that might be
Characterization
responsible for dye binding (D’Souza et al., 2008). There are no
X-ray diffraction (XRD) traces of halloysite were acquired using a significant differences among the spectra before and after adsorp-
PANalytical X’Pert PRO powder diffractometer in θ–θ configur- tion. Therefore, the functional groups (hydroxyls and aluminosili-
ation with an X’Celerator detector and variable divergence and cate framework) noted in the FTIR spectra may not have been
fixed receiving slits with Fe-filtered Co-Kα radiation (λ = 1.789 Å). involved in the dye-removal process.
The phases were identified using X’Pert Highscore Plus software.
The relative phase amounts (wt.%) were estimated using the TEM results
Rietveld method (Autoquan program). Fourier-transform infrared The TEM images of the original halloysite and the halloysite
(FTIR) spectroscopy was used to determine the functional groups after adsorption of DR81, MO, MB and CV dyes are shown in
of halloysite. A powdered sample of the adsorbent material was Fig. 3. The original halloysite consists mainly of halloysite nano-
analyzed using ALPHA II’s Platinum attenuated total reflection tubes, some of which exhibit hollow, rod-like structures
(ATR) single reflection diamond module with the ATR method of (Joussein et al., 2005; Ahmed et al., 2010). The halloysite
Clay Minerals 199

Fig. 2. FTIR spectra of the original halloysite and the halloysite after adsorption of
DR81, MO, MB and CV dyes.
Fig. 1. XRD traces of the original halloysite and the halloysite after adsorption of
DR81, MO, MB and CV dyes.

HMO. A large specific surface area is associated with significant


nanotubes are 500–700 nm long, and their inner and outer dia- soil–water contaminant interaction, which indicates high reactiv-
meters are 40–60 nm. The nanotubes contain empty lumens, ity (Pennel, 2016). Theoretical relationships for surfaces of various
with some rod structures being inside the empty lumens particle geometries show that surface area is controlled by the
(Fig. 3a,b). The nanotubes of halloysite after adsorption of dyes smallest particle size. With decreasing particle size, the specific
contain additional material on the external walls (Fig. 3c–f). surface area increases and the small particles tend to have platy
Moreover, the halloysite maintains its tubular structure even and rod-like geometries (Santamarina et al., 2002). After adsorp-
after dye adsorption. tion of the four dyes, there is a clear change in the shape of the
particles (Fig. 4), particularly with the adsorption of DR81 show-
ing platy geometries; hence, this might have increased the specific
SEM results
surface area after dye adsorption. The kaolinite pores are more
The SEM images of the original halloysite and the halloysite after
open, and this leads to a greater pore diameter of the slit-shaped
adsorption of DR81, MO, MB and CV dyes are shown in Fig. 4.
pores in the structure (Boukhemkhem & Rida, 2017). As
The halloysite consists of agglomerates of nanotubes with some
explained previously, the fibrous surface topography of the clays
irregularities in their diameters and wall thicknesses and a fibrous
is expected to increase the degree of unsaturation of the surface
morphology (Kamble et al., 2012). The original halloysite has an
with dangled bonds and free valences, thus increasing the total
uneven surface, which nevertheless is smoother compared to its
surface area of the clay (Ngulube et al., 2017).
counterpart after dye adsorption. After dye adsorption, the surface
morphology changed considerably, which indicates that dye mole- Batch dye adsorption studies
cules were adsorbed onto the surface and interaggregate pores,
because the halloysite surface is more uniform, especially after The effect of contact time, adsorbent dosage, initial concentration,
adsorption of MB and CV dyes (Fig. 4e,f). Previous studies also pH, temperature and PZC on the adsorption of dyes is shown in
revealed that the surface of the adsorbent became more homogen- Fig. 5.
ous compared to the original materials (Sarma et al., 2016). In add-
ition, the halloysite has a different morphology after adsorption of Effect of time
MO and CV (Fig. 4d,f). The images show the rod-like structures of The kinetics of the adsorption determine the optimum time at
the halloysite, clearly confirming the tubular nature of the clay. which equilibrium is attained. Depending on wastewater-process
Figure 4d shows a homogenous surface consisting of fluffy flakes design, this information may be used in wastewater-treatment
of various sizes and shapes. The fibrous surface topography of the plants. The results of kinetic tests are shown in Fig. 5a. The
clays is expected to increase the degree of unsaturation of the surface adsorption of all dyes was comparable for all of the contact
with dangled bonds and free valences, thus increasing the total spe- times used. The removal of MB and CV was almost complete
cific surface area of the clay (Ngulube et al., 2017). (98–99%), that of DR81 was slightly lower (86–87%) and finally
that of MO was the lowest (79–80%). The near-constant percent-
BET results age of removal after 30 min indicates that the adsorbent was
The large adsorption capacities of clay materials are largely due to approaching its optimum performance; hence, 30 min was
their large specific surface areas (Cadena et al., 1990). The specific selected as the optimum contact time for subsequent experiments.
surface area of a clay is directly proportional to its adsorption cap-
acity (Santos et al., 2016; Sarma et al., 2016). The halloysite has a Effect of adsorbent dose
BET specific surface area of 41.94 m2/g (Table 1), which is greater The effect of adsorbent dosage is illustrated in Fig. 5b. With
than or comparable to other clays (Park et al., 2011; Gu et al., increasing adsorbent dose, the percentage of removal of the
2014; Sari & Tuzen, 2014). Comparison between the original hal- dyes increased, similarly to previous studies (Ciesielczyk et al.,
loysite (H) and its counterparts after dye adsorption (HDR81, 2017). Increasing adsorbent dosage at a constant volume and
HMO, HMB and HCV) showed an increase in the specific surface dye concentration leads to the saturation of adsorption sites.
area after dye adsorption, with the greatest surface area being for Due to the formation of aggregates at higher doses, there might
200 Tholiso Ngulube et al.

Fig. 3. TEM images of the original halloysite (a,b) and the halloy-
site after adsorption of (c) DR81, (d) MO, (e) MB and (f) CV dyes.

be a decrease in adsorption capacity because aggregation would Effect of pH


lead to a decrease in total surface area of the adsorbent and The pH affects the adsorption of dyes because it affects the surface
increased diffusion path length (Sarma et al., 2016). characteristics of the adsorbent (Ngulube et al., 2018). The pH of
a solution determines the degree of protonation of exchange sites
Effect of initial dye concentration of the adsorbate due to variable charge, thereby influencing the
The effect of initial dye concentration is shown in Fig. 5c. By specific charge of exchange sites and eventually the adsorption
increasing initial dye concentration to 20 mg/L, the percentage tendency of the adsorbent (Tomar et al., 2014). In the present
of removal of DR81 and MO increased, decreasing at higher con- study, the pH had a negligible influence on the removal of the
centrations. However, the percentage of removal of CV and MB DR81, MB and CV dyes in the pH range 2–12 (Fig. 5d), in
showed a different trend, decreasing gradually with increasing ini- accordance with previous studies (Santos et al., 2016). For MO,
tial dye concentration from 1 to 40 mg/L. This observation is in maximum removal (70.24%) was observed at pH 8, being lower
accord with previous studies (e.g. Anirudhan & Ramachandran, than 40% for the remaining pH values. The MO molecules have
2015) in which the concentration of the dye was a driving force different charges at different pH values (Subbaiah & Kim,
for overcoming the mass transfer resistance of dye molecules 2016). The dissociation constant (pKa) for MO is 3.46 (Arshadi
between the solid and the aqueous phases. At low dye concentra- et al., 2016); hence, the MO species will mainly exist as monova-
tions, there are more active sites available for adsorption; hence, at lent anions at this equilibrium pH. Below and above this pH, the
this stage, adsorption is independent of initial concentration. MO will go through a protonation process that will determine its
However, with increasing concentration, the available sites are charge. Due to the protonation reaction (≡RN(CH3)2 + H+ →
occupied and ultimately a decrease in dye removal is observed ≡RNH+(CH3)2), MO is positively charged at pH < 3.76 (Yao
(Ogunmodede et al., 2015). et al., 2011). In addition, the halloysite surface is also positively
Clay Minerals 201

Fig. 4. SEM images of the original halloysite (a,b) and the halloy-
site after adsorption of (c) DR81, (d) MO, (e) MB and (f) CV dyes.

Table 1. Specific surface areas of the original halloysite (H) and the halloysite adsorption. An increase in adsorption capacity with temperature
after adsorption of DR81 (HDR81), MO (HMO), MB (HMB) and CV (HCV) dyes. indicates an endothermic reaction whereas a decrease in adsorp-
tion capacity with increasing temperature indicates an exothermic
BET surface
Single point surface area area Langmuir surface area process (Zulfikar, 2013). In the present study, temperature varia-
(m2/g) (m2/g) (m2/g) tions did not affect significantly the removal of all four dyes, indi-
cating that there is no need to adjust temperature in order to
Halloysite 41.0 41.9 57.4
effectively remove dyes from aqueous solutions using halloysite
HDR81 44.7 45.8 62.8
HMO 45.5 46.6 65.0 (Fig. 5f).
HMB 42.5 43.5 61.0
HCV 42.9 44.0 60.3 Point of zero charge studies
The PZC determines the type of surface-active centres on the
adsorbent, which affect the adsorption processes (Yagub et al.,
2014). Adsorption of positively charged dyes is enhanced at
charged; hence, adsorption is limited because of repulsive forces
pH > PZC due to the presence of functional groups such as
between the surface and the positively charged MO particles.
hydroxyl ions on the surface, whereas adsorption of negatively
With increasing pH, the MO has fewer positive charges. Hence,
charged dyes is enhanced at pH < PZC, because the surface of
the enhanced MO adsorption on halloysite is attributed to hydro-
the adsorbent is positively charged (Ngulube et al., 2017).
gen bonding, electrostatic attraction and strong surface complex-
When the pH is higher than the PZC, the surface charge of hal-
ation (Xiang et al., 2014). At alkaline pH up to 12, the electrostatic
loysite is negative, while it is positive when the pH is lower than
repulsion and OH– competition lead to a decrease in the adsorp-
the PZC. The PZC of the halloysite used in this study is 5
tion of anions (Cheah et al., 2013).
(Fig. 5f). It would be expected that the adsorption of MB and CV
would be greatest at pH > 5 and those of DR81 and MO would be
Effect of temperature greatest at pH < 5. However, the variation of pH did not affect the
Generally, temperature has a significant effect on reaction rate. removal of the dyes, except for MO (Fig. 5d), suggesting that the
With increasing temperature, chemical reaction rate also increases adsorption of the remaining three dyes was not predominantly
(Mekatel et al., 2015). Therefore, temperature might affect dependent on the surface charges of halloysite.
202 Tholiso Ngulube et al.

Fig. 5. Removal of DR81, MO, MB and CV dyes by halloysite as a function of (a) contact time, (b) halloysite dosage, (c) initial dye concentration, (d) pH, (e) tem-
perature and (f) PZC.

Treatment of real industrial effluent Table 2. Characterization of actual industrial effluent.

The physical characterization of textile mill raw wastewater has Industrial effluent Industrial effluent
been carried out for colour, total dissolved solids, electrical con- before treatment after treatment
ductivity, turbidity, total suspended solids and pH. The results Analyser parameter with halloysite with halloysite
of the wastewater characteristics before and after treatment with Total dissolved solids (mg/L) 1219 1068
halloysite nanoclay are presented in Table 2 and the progress of Total suspended solids (mg/L) 918 785
the reaction is illustrated in Fig. 6. The wastewater is classified Electrical conductivity (μS/m) 1764 1504
as ‘high-strength wastewater’ (Benefield, 2002). After treatment Turbidity (NTU, FNU) 0.01 246
with halloysite clay at the optimum conditions determined from pH 8.77 8.20
Colour Black Milky white
the batch study, the wastewater could be characterized as medium-
strength wastewater. The dye-removal rates were 79.25%, 90.01%, FNU = Formazin Nephelometric Unit; NTU = Nephelometric Turbidity Unit.
Clay Minerals 203

Table 3. The pseudo-second-order parameters for the adsorption of DR81, MO,


MB and CV dyes onto the halloysite clay.

Parameter DR81 MO MB CV

Qe (mg/g) experimental 4.3553 0.2055 4.9219 1.2419


Qe (mg/g) calculated 4.3668 4.7847 4.9285 5.2192
k2 (g/mg/h) 0.0001 1154.9 6.0725 630.56
R2 1 1 1 1

The fitting between experimental data and the model-predicted


values was expressed by the coefficient of determination, R2. The
curves of the plots are given in the Supplementary Material, and
the calculated kinetic parameters and the corresponding linear
regression correlation coefficient R2 values are summarized in
Table 3. There is good agreement between experimental and
calculated qe values for DR81 and MB dyes. The coefficients of
determination for the pseudo-second-order kinetics model (R2)
are all equal to 1, indicating the applicability of this kinetic model
and the second-order nature of the adsorption process of the
dyes onto halloysite clay. These results imply that the chemisorp-
tion mechanism may play an important role in the adsorption of
DR81, MO, MB and CV dyes.

Adsorption isotherm studies


An adsorption isotherm describes the relationship between the
Fig. 6. Real industrial effluent before and after treatment with halloysite nanoclay. amount of adsorbate adsorbed by the adsorbent and the remain-
ing adsorbate concentration in the solution at equilibrium.
Adsorption equilibrium studies were performed using two
92.22% and 92.22% for DR81, MO, MB and CV dyes, respect- adsorption isotherms.
ively. The colour of the textile wastewater changed from black The Langmuir isotherm describes the formation of a mono-
to milky white after treatment (Fig. 6). layer adsorbate on the outer surface of the adsorbent, beyond
which no further adsorption takes place (Langmuir, 1918). The
isotherm is expressed by the following equation:
Adsorption kinetics studies
 
Adsorption kinetics are important in determining the effective- Ce 1 1
= Ce + (3)
ness of adsorption. In this study, two models were used to predict qe Qm Qm b
the adsorption kinetics of DR81, MO, MB and CV onto halloysite.
where Ce is the equilibrium concentration (mg/L), Qe is the
amount adsorbed at equilibrium (mg/g), b represents the
The pseudo-first-order model Langmuir isotherm constant (L/mg) and Qm is the maximum
The linear form of the pseudo-first-order model was described by adsorption capacity (mg/g) for complete monolayer coverage.
Ho & McKay (1998) and it can be expressed as: The separation factor (RL) was calculated from the following
  equation:
k1 t
log(qe − qt ) = log(qe ) − (1) 1
2.303 RL = (4)
(1 + bCo )
Where qe (mg/g) is the adsorption capacity at equilibrium, qt
(mg/g) is the adsorption capacity at time t and k1 (1/min) is where Co is the initial concentration (mg/L). RL values between 0
the rate constant. and 1 indicate favourable adsorption (0 indicates irreversible
adsorption, 1 suggests linear adsorption), while a value greater
than 1 indicates unfavourable adsorption.
Pseudo-second-order model The Freundlich isotherm is commonly used to describe het-
The pseudo-second-order model, which has been applied for ana- erogeneous systems (Freundlich, 1906). The model is represented
lysing chemisorption kinetics on solid phases from liquid solu- in linear form via the following equation:
tions (Ho, 2006), is expressed linearly as:
    logQe = nlogCe + logKF (5)
t 1 1
= t+ (2)
qt qe k 2 q2e where KF (mg/g) and n are the Freundlich constants describing
the adsorption capacity and intensity, respectively. The constants
where k2 is the rate constant (g/mg/h) and k2 q2e (mg/g/h) is the n and KF are determined from the slope and the intercept,
initial adsorption rate. respectively, of the plot logQe versus logCe. Values of n between
204 Tholiso Ngulube et al.

Table 4. Parameters of the Langmuir adsorption isotherms for the adsorption Table 5. Thermodynamic parameters for the adsorption of DR81, MO, MB and
of DR81, MO, MB and CV dyes onto the halloysite clay. CV dyes onto the halloysite clay.

Parameters DR81 MO MB CV Parameters DR81 MO MB CV

Langmuir isotherm ΔG° (kJ/mol) –15,341.01 –7242.34 –4560.09 –12,917.68


Qm (mg/g) 67.57 104.17 185.16 5.22 ΔH° (kJ/mol) –25,776.73 3745.21 547.29 219.57
Ka (L/mg) 85.25 110.82 119.45 12.08 ΔS° (kJ/mol/K) –5.22 4.59 23.22 4.23
RL 0.3719 0.2775 0.1777 0.8846 Ka (L/g) 331,189.87 4.57 1.26 1.09
R2 0.0083 0.8033 0.9969 0.9784
Freundlich isotherm
R2 0.6589 0.6321 0.7975 0.8093
The adsorption occurs through the formation of an activated
n 0.8599 1.2097 0.1831 1.8278
KF (mg/g) 1.7037 0.0428 12.4824 4.3642 complex between the adsorbate and adsorbent, thus representing
the activation stage at that level of reaction. Spontaneous adsorp-
tion of chemical species on stevensite was also reported by Hajjaji
et al. (2018). However, in the case of MO, MB and CV, the
1 and 10 indicate favourable adsorption (Kadirvelu & Namasivayam, entropy change values were positive, indicating that there was a
2000). change in the internal structure of the halloysite during adsorp-
The equation parameters of the equilibrium models give insight tion (Ofomaja et al., 2010).
into the adsorption mechanism, surface properties and affinity of
the adsorbent for the adsorbates. The Langmuir isotherm yielded
high coefficients of determination for MO, MB and CV dyes, Regeneration studies
this being the best-fitting model for adsorption of the dyes onto The results of the regeneration of the halloysite clay for the
halloysite (Table 4). In contrast, the Freundlich isotherm yielded removal of acidic and basic dyes from aqueous solutions are
low correlation coefficients for all of the dyes, hence the parameters shown in Fig. 7. The spent adsorbent has less adsorption capacity
of the Langmuir isotherm were used to describe the mechanism of than the original adsorbent. The halloysite displayed a similar
adsorption. In accordance with this study, Hajjaji et al. (2016) used trend after adsorption of MB and DR81 dyes, whereas this
the Langmuir isotherm model parameters to fit their results, sug- trend was different after adsorption of CV and MO dyes. After
gesting that the adsorption corresponds to monolayer coverage of one adsorption cycle with MB and DR81 dyes, the spent adsorb-
MB molecules over the kaolinite surface. The dimensionless equi- ent lost its capacity to remove dyes on successive cycles. By con-
librium separation factor, RL, for all of the dyes was <1, indicating trast, after one adsorption cycle, the adsorption capacity for CV
that dye adsorption was a favourable process. and MO dyes decreased after the first regeneration cycle, but
for successive cycles, the dye removal increased as the halloysite
was regenerated. This behaviour indicates that the adsorption of
Thermodynamic study
CV and MO dyes is chemical in nature. The adsorption was
The enthalpy ΔH° (kJ/mol) and entropy ΔS° (J/mol/K) of the reversible and the adsorbed CV and MO dyes were almost com-
adsorption were determined using Eqs 6 and 7. pletely recovered with the alkaline solution. A probable explan-
The standard free energy change during adsorption, ΔG°, was ation for this behaviour would be that MO is an acidic dye
calculated from the following equation: causing electrostatic interactions with the alkaline solution and
hence reversible reactions. CV is a water-soluble organic com-
DG◦ = −RTlnKa (6) pound with a violet cation. Adding NaOH converts the cation
into a carbonyl base, and the polarity of oxygen also makes the
where ΔG is the free energy of sorption (kJ/mol), T is the tem- α-hydrogens of carbonyl compounds much more acidic, causing
perature in K, R is the universal gas constant (8.314 J/mol/K) possible acid–base electrostatic attractions (Cannon et al., 1994).
and Ka is the sorption equilibrium constant.
The sorption equilibrium constant Ka can be expressed in Dye-adsorption mechanisms
terms of enthalpy change (ΔH) and entropy change (ΔS) as a
function of temperature: The adsorption studies and characterization might allow the prop-
osition of mechanisms for the adsorption of the anionic and cat-
DH ◦ DS◦ ionic dyes on the halloysite clay. For clays, it is expected that the
lnKa = + (7) solution pH and ultimately the PZC influence the adsorption of
RT RT
anionic and cationic molecules depending on the extent of proton-
where ΔH° is the heat of sorption (kJ/mol) and ΔS° is the standard ation of the adsorbate and the adsorbent. According to the PZC
entropy change (kJ/mol/K). results, it is anticipated that the adsorption of MB and CV would
The calculated values for the thermodynamic parameters are be greatest at pH > 5 and that of DR81 and MO would be greatest
listed in Table 5. The enthalpy of activation, ΔH°, is positive for at pH < 5. However, Fig. 5d shows that the variation of pH had a
all of the dyes except for DR81, indicating that the adsorption negligible influence on the percentage of removal of the dyes, except
of MO, MB and CV dyes on the halloysite clay is endothermic, for MO. This might suggest that the solution pH did not cause not-
whereas that of DR81 dye is exothermic. The adsorption of all able protonation of the ions in solution and subsequently the impact
the dyes was spontaneous, as symbolized by the negative free of pH on the dye uptake process was negligible. Similarly the FTIR
energy of sorption values (–ΔG°). The negative value of entropy spectra did not show any notable activity or change of the functional
change, ΔS°, for DR81 indicates that no significant changes groups before and after adsorption of the dyes (Fig. 2).
occurred at the activation stage. A negative value of ΔS° suggests Nonetheless, the overall observation was that, of the four dyes,
that the adsorption process involves an associative mechanism. MO had the lowest percentage of dye removal. The molecular size
Clay Minerals 205

Fig. 7. Regeneration of the halloysite clay.

Table 6. Adsorption capacities for various adsorbent materials used to remove dyes.

Adsorbent Adsorption capacity (mg/g) Dye Reference

Halloysite clay 104.17 Methyl Orange Present study


Halloysite clay 185.16 Methylene Blue Present study
Halloysite clay 5.22 Crystal Violet Present study
Halloysite clay 67.57 Direct Red 81 Present study
HDTMA-modified halloysites 91.74 Methyl Orange Liu et al. (2012)
Halloysite nanotubes 54.85 Neutral Red Luo et al. (2010)
Halloysite nanotubes 113.64 Methyl Violet Liu et al. (2011)
Chitosan–halloysite composite 68.92 Methylene Blue Peng et al. (2015)
Magnetic halloysite nanotubules 37.38 Methylene Blue Xie et al. (2011)
Magnetic halloysite nanotubules 31.21 Nile Red Xie et al. (2011)
Magnetic halloysite nanotubules 0.68 Methyl Orange Xie et al. (2011)

HDTMA = hexadecyltrimethylammonium.

of the adsorbates also affects adsorption. Large molecules might not Process 2: In basic pH conditions and at the particle edges of disso-
be adsorbed because their size prevents their diffusion into the pores ciated aluminol groups, OH groups may be substituted by anions
of the adsorbent. The chemical formulae and molecular weights of (Errais et al., 2012).
Process 3: Halloysite is a kaolinite clay and these clays have a net layer
DR81, MB, CV and MO are C29H19N2O8S2 (675.598 g/mol),
charge of zero (Sarma et al., 2011). However, a small net negative charge
C16H18N3ClS.xH2O (319.850 g/mol), C25N3H30Cl (407.979 g/mol)
may arise from broken edges on the clay crystals. This negative charge may
and C14H14N3NaO3S (327.334 g/mol), respectively. Hence, MO be responsible for holding the cations to the clay surface through coulom-
has the second smallest molecule, and if molecular size had affected bic interactions. The charge on the edges is due to protonation or depro-
the adsorption of the dyes, MO would have presented a greater per- tonation of surface hydroxyl groups (Kausar et al., 2018).
centage of removal compared to MB and DR81.
As the dye uptake could not be explained in terms of molecu- In general, the uptake of dye molecules by clay minerals is due to
lar size, pH variation or characterization results, the adsorption the trapping of dye molecules by exchangeable ions and the pro-
mechanism will have to be discussed in light of the isotherm tonated edges of the clay layers.
and kinetic modelling results. The pseudo-second-order model
showed the best fit with the experimental data for all four of Comparison with other studies
the dyes, whereas the Langmuir isotherm model described accur-
Table 6 illustrates the comparison of the adsorption capacity of hal-
ately the adsorptions of MB, CV and MO, but not that of DR81.
loysite with other clay-based adsorbents. The adsorption capacity of
Hence, the adsorption mechanisms may involve complexation or
the halloysite clay used for this study for MO, MB and DR81 dyes
ion exchange between the dye ions and the surface ions or SiO2 or
is greater than the capacities of other halloysite clays, whereas the
Al2O3 groups on the clay surfaces. The halloysite clay displayed
capacity for the removal of CV dye is notably lower. This difference
relatively high Langmuir adsorption capacities for all of the dyes
might be attributed to the various operational parameters used dur-
and therefore most of the dye ions might have been held in a
ing the adsorption experiments and to the various chemical mod-
monolayer on the clay surface.
ifications performed on the clay before use. Thus, the halloysite
According to Errais et al. (2012), depending on the type of
used may be considered as an effective option for the removal of
clay, the possible processes involved in dye adsorption can be
both acidic and basic dyes from aqueous media.
summarized in the following processes:

Process 1: The polyvalent metal cations (Si4+ and Al3+) found on edge Conclusion
sites of the clay contribute to the surface charge being positive, hence
forming bridges between the particles of the clay and organic anions, The removal of DR81, MO, MB and CV dyes from aqueous solu-
allowing anionic dyes such as DR81 to be adsorbed. tion was carried out using a halloysite clay as an adsorbent. The
206 Tholiso Ngulube et al.

optimum contact time for the interaction of the adsorbent with Gamoudi S. & Srasra E. (2018) Removal of cationic and anionic dyes using
dye effluent was 30 min. The pseudo-second-order kinetic purified and surfactant-modified Tunisian clays: kinetic, isotherm, thermo-
model described sufficiently the adsorption, yielding a perfect dynamic and adsorption-mechanism studies. Clay Minerals, 53, 159–174.
Gu Z., Gao M., Luo Z., Lu L., Ye Y. & Liu Y. (2014) Bis-pyridinium dibromides
coefficient of determination (R2 = 1) for all of the dyes. The
modified organobentonite for the removal of aniline from wastewater: a posi-
Langmuir model described the adsorption process better than
tive role of π–π polar interaction. Applied Surface Science, 290, 107–115.
did the Freundlich model in the order MB > CV > MO > Hajjaji M., Beraa A., Coppel Y., Laurent R. & Caminade A. (2018) Adsorption cap-
DR81 in terms of the R2 values. The untreated halloysite clay is acity of sodic- and dendrimers-modified stevensite. Clay Minerals, 53, 525–544.
an effective material for adsorbing acidic and basic dyes at a nor- Hajjaji W., Andrejkovičová S., Pullar R., Tobaldi D., Lopez-Galindo A.,
mal temperature and pH of solution. The results have established Jammousi F. & Labrincha J. (2016) Effective removal of anionic and cationic
the potential for the halloysite clay to be used as an adsorbent for dyes by kaolinite and TiO2/kaolinite composites. Clay Minerals, 51, 19–27.
the removal of dyes from wastewater in the textile industry. Ho Y.S. & McKay G. (1998) A comparison of chemisorption kinetic models
applied to pollutant removal on various sorbents. Process Safety and
Supplementary material. To view supplementary material for this article, Environmental Protection, 76, 332–340.
please visit https://doi.org/10.1180/clm.2019.27 Ho Y.S. (2006) Review of second-order models for adsorption systems. Journal
of Hazardous Materials B, 136, 681–689.
Author ORCIDs. T. Ngulube, 0000-0003-0725-2203. Izuagie A.A., Gitari W.M. & Gumbo J.R. (2016) Synthesis and performance
evaluation of Al/Fe oxide coated diatomaceous earth in groundwater
Acknowledgements. The authors thank the University of Venda, National
defluoridation: towards fluorosis mitigation. Journal of Environmental
Research Foundation (NRF), Water Research Commission (WRC) for funding.
Science and Health A, 51, 810–824.
Jiang L., Huang Y. & Liu T. (2015) Enhanced visible-light photocatalytic per-
formance of electrospun carbon-doped TiO2/halloysite nanotube hybrid
References
nanofibers. Journal of Colloid and Interface Science, 439, 62–68.
Adebowale K.O., Olu-Owolabi B.I. & Chigbundu E.C. (2014) Removal of Joussein E., Petit S., Churchman J., Theng B., Righi D. & Delvaux B. (2005)
safranin-O from aqueous solution by adsorption onto kaolinite clay. Halloysite clay minerals – a review. Clay Minerals, 40, 383–426.
Journal of Encapsulation and Adsorption Sciences, 4, 89–104. Kadirvelu K. & Namasivayam C. (2000) Agricultural by-product as metal
Ahmed S., Rasul M.G., Martens W.N., Brown R. & Hashib M.A. (2010) adsorbent: sorption of lead (II) from aqueous solution onto coirpith carbon.
Heterogeneous photocatalytic degradation of phenols in wastewater: a Environmental Technology, 21, 1091–1097.
review on current status and developments. Desalination, 261, 3–18. Kamble R., Ghag M., Gaikawad S. & Panda B.J. (2012) Halloysite nanotubes
Anirudhan T.S. & Ramachandran M. (2015) Adsorptive removal of basic dyes and applications: a review. Journal of Advanced Science Research, 3, 25–29.
from aqueous solutions by surfactant modified halloysite clay (organoclay): Kausar A., Iqbal M., Javeda A., Aftab K., Nazli Z., Bhatti H.N. & Nouren S.
kinetic and competitive adsorption isotherm. Process Safety and (2018) Dyes adsorption using clay and modified clay: a review. Journal of
Environmental Protection, 95, 215–225. Molecular Liquids, 256, 395–407.
Arshadi M., Mousavini F., Amiri M.J. & Faraji A.R. (2016) Adsorption of Langmuir I. (1918) The adsorption of gases on plane surfaces of glass, mica
methyl orange and salicylic acid on a nano-transition metal composite: and platinum. Journal of the American Chemical Society, 40, 1361–1403.
kinetics, thermodynamic and electrochemical studies. Journal of Colloid Liu R., Fu K., Zhang B., Mei D., Zhang H. & Liu J. (2012) Removal of Methyl
and Interface Science, 483, 118–131. Orange by modified halloysite nanotubes. Journal of Dispersion Science and
Benefield L.A. (2002) Wastewater Quality/Strength/Content. Wastewater Technology, 33, 711–718.
Management Program. Washington State Department of Health, Liu R., Zhang B., Mei D., Zhang H. & Liu J. (2011) Adsorption of Methyl
Turnwater, WA, USA. Violet from aqueous solution by halloysite nanotubes. Desalination, 268,
Berthier P. (1826) Analyse de l’halloysite. Annales de Chimie et de Physique, 111–116.
32, 332–335. Luo P., Zhao Y.F., Zhang B., Liu J.D., Yang Y. & Liu J.F. (2010) Study on the
Boukhemkhem A. & Rida K. (2017) Improvement adsorption capacity of adsorption of Neutral Red from aqueous solution onto halloysite nanotubes.
methylene blue onto modified Tamazert kaolin. Adsorption Science & Water Research, 44, 1489–1497.
Technology, 35, 752–773. Mekatel E., Amokrane S., Aid A., Nibou D. & Trari M. (2015) Adsorption of
Cadena F., Rizvi R. & Peters R.W. (1990) Feasibility studies for the removal of methyl orange on nanoparticles of a synthetic zeolite NaA/CuO. Comptes
heavy metals from solution using tailored bentonite. Presented at: Rendus Chimie, 18, 336–344.
Hazardous and Industrial Wastes Twenty-Second Mid-Atlantic Industrial Mu B. & Wang A. (2016) Adsorption of dyes onto palygorskite and its com-
Waste Conference. Drexel University, Philadelphia, PA, USA. posites: a review. Journal of Environmental Chemical Engineering, 4, 1274–
Cannon J.F., Gammon S.D. & Hunsberger L.R. (1994) KineticsLab: the Crystal 1294.
Violet/sodium hydroxide reaction. Journal of Chemical Education, 71, 238. Ngulube T., Gumbo J.R., Masindi V. & Maity A. (2017) An update on syn-
Cheah W., Hosseini S., Khan M.A., Chuah T.G. & Choong T.S.Y. (2013) thetic dyes adsorption onto clay-based minerals: a state-of-art review.
Carbon coated monolith, a mesoporous material for the removal of methyl Journal of Environmental Management, 191, 35–57.
orange from aqueous phase: adsorption and desorption studies. Chemical Ngulube T., Gumbo J.R., Masindi V. & Maity A. (2018) Calcined magnesite as
Engineering Journal, 215–216, 747–754. an adsorbent for cationic and anionic dyes: characterization, adsorption
Ciesielczyk F., Bartczak P., Zdarta J. & Jesionowski T. (2017) Active MgO–SiO2 parameters, isotherms and kinetics study. Heliyon, 4, e00838.
hybrid material for organic dye removal: a mechanism and interaction study Ofomaja A.E. (2010) Intraparticle diffusion process for lead (II) biosorption
of the adsorption of C.I. Acid Blue 29 and C.I. Basic Blue 9. Journal of onto mansonia wood sawdust. Bioresource Technology, 101, 5868–5876.
Environmental Management, 204, 123–135. Ogunmodede O.T., Ojo A.A., Adewole E. & Adebayo O.L. (2015) Adsorptive
D’Souza L., Devi P., Divya S.M.P. & Naik C.G. (2008) Use of Fourier transform removal of anionic dye from aqueous solutions by mixture of kaolin and
infrared (FTIR) spectroscopy to study cadmium-induced changes in Padina bentonite clay: characteristics, isotherm, kinetic and thermodynamic stud-
tetrastromatica (Hauck). Analytical Chemical Insights, 3, 135–143. ies. Iranian Journal of Energy and Environment, 6, 147–153.
Errais E., Duplaya J., Elhabiri M., Khodja M., Ocampo R., Baltenweck- Park Y., Ayoko G.A. & Frost R.L. (2011) Characterisation of organoclays and
Guyot R. & Darragi F., (2012) Anionic RR120 dye adsorption onto raw adsorption of p-nitrophenol: environmental application. Journal of Colloid
clay: surface properties and adsorption mechanism. Colloids and Surfaces Interface Science, 360, 440–456.
A: Physicochemical Engineering Aspects, 403, 69–78. Peng Q., Liu M., Zheng J. & Zhou C. (2015) Adsorption of dyes in aqueous
Freundlich H.Z. (1906) Over the adsorption in solution. Journal of Physical solutions by chitosan–halloysite nanotubes composite hydrogel beads.
Chemistry, 57A, 385–470. Microporous and Mesoporous Materials, 201, 190–201.
Clay Minerals 207

Pennell K.D. (2016) Specific surface area. Pp. 1–8 in: Reference Module in Tomar V., Prasad S. & Kumar D. (2014) Adsorptive removal of fluoride from
Earth Systems and Environmental Sciences Publisher (A.E. Scott, editor). aqueous media using citrus limonum (lemon) leaf. Microchemal Journal,
Elsevier, New York, NY, USA. 112, 97–103.
Prokop A., Maziarz P. & Matusik J. (2015) Removal of selected anions by Xiang W., Zhang G.K., Zhang Y.L., Tang D.D. & Wang J.T. (2014) Synthesis
raw halloysite and smectite clay. Geology, Geophysics and Environment, and characterization of cotton-like Ca–Al–La composite as an adsorbent for
41, 125–126. fluoride removal. Chemical Engineering Journal, 250, 423–430.
Sari A. & Tuzen M. (2014) Cd (II) adsorption from aqueous solution by raw Xie Y., Qian D., Wu D. & Ma X. (2011) Magnetic halloysite nanotubes/iron
and modified kaolinite. Applied Clay Science, 88–89, 63–72. oxide composites for the adsorption of dyes. Chemical Engineering
Sarma G.K., Gupta S.S. & Bhattacharyya K.G. (2016) Adsorption of Crystal Journal, 168, 959–963.
Violet on raw and acid-treated montmorillonite, K10, in aqueous suspen- Yagub M.T., Sen T.K., Afroze S. & Ang H.M. (2014) Dye and its removal from
sion. Journal of Environmental Management, 171, 1–10. aqueous solution by adsorption: a review. Advanced Colloid Interface
Sarma G.K., Gupta S.S. & Bhattacharyya K.G. (2011) Methylene Blue adsorp- Science, 209, 172–184.
tion on natural and modified clays. Separation Science and Technology, 46, Yang R., Li D., Li A. & Yang H. (2018) Adsorption properties and mechanisms
1602–1614. of palygorskite for removal of various ionic dyes from water. Applied Clay
Santamarina J.C., Klein K.A., Wang Y. & Prencke E. (2002) Specific surface: Science, 151, 20–28.
determination and relevance. Canadian Geotechnical Journal, 39, 233–241. Yao Y., Bing H., Feifei X. & Xiaofeng C. (2011) Equilibrium and kinetic studies
Santos S.C.R., Oliveira A.F.M. & Boaventura R.A.R. (2016) Bentonitic clay as of Methyl Orange adsorption on multiwalled carbon nanotubes. Chemical
adsorbent for the decolourisation of dyehouse effluents. Journal of Engineering Journal, 170, 82–89.
Cleaner Production, 126, 667–676. Zhang Y., Fua L. & Yang H. (2012) Insights into the physicochemical aspects
Słomkiewicz P.M., Szczepanik B. & Garnuszek M. (2015) Determination of from natural halloysite to silica nanotubes. Colloids and Surfaces A:
adsorption isotherms of aniline and 4-chloroaniline on halloysite Physicochemical Engineering Aspects, 414, 115–119.
adsorbent by inverse liquid chromatography. Applied Clay Science, 114, Zhao Y., Abdullayev E., Vasiliv A. & Lvov Y. (2013) Halloysite nanotubule clay
221–228. for efficient water purification. Journal of Colloid and Interface Science, 406,
Subbaiah M.V. & Kim D.S. (2016) Adsorption of Methyl Orange from aqueous 121–129.
solution by aminated pumpkin seed powder: kinetics, isotherms, and Zulfikar M.A. (2013) Materials, effect of temperature on adsorption of humic
thermodynamic studies. Ecotoxicology and Environmental Safety, 128, acid from peat water onto pyrophyllite. International Journal of Chemical
109–117. Environmental and Biological Sciences, 1, 88–91.
Copyright © Mineralogical Society of Great Britain and Ireland 2019

You might also like