You are on page 1of 13

Journal of Cleaner Production 207 (2019) 350e362

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

Synthesis of porous biomass fly ash-based geopolymer spheres for


efficient removal of methylene blue from wastewaters
Rui M. Novais*, Joa
~o Carvalheiras, David M. Tobaldi, Maria P. Seabra, Robert C. Pullar,
~o A. Labrincha
Joa
rio de Santiago, 3810-193
Department of Materials and Ceramic Engineering / CICECO- Aveiro Institute of Materials, University of Aveiro, Campus Universita
Aveiro, Portugal

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

Article history: In this work, and for the first time, fly ash (FA)-based geopolymer (d ¼ 2.6 mm) spheres were used to
Received 10 April 2018 extract methylene blue from synthetic wastewaters. The influence of sorption time, dye initial concen-
Received in revised form tration and adsorbent amount on the dye removal efficiency and uptake by the porous spheres was
12 September 2018
evaluated. The adsorbents' recyclability and their dye fixation efficiency were also considered. The initial
Accepted 30 September 2018
Available online 3 October 2018
dye concentration strongly affected the uptake and removal efficiency by the porous bodies, the former
rising from 1.1 to 30.1 mg/g when the dye initial concentration jumped from 10 to 250 ppm, and the
latter increasing from 82.3% to 94.3% when the dye initial concentration varied from 10 to 125 ppm.
Keywords:
Adsorption
Results showed a much faster (24 h) and higher (30.1 mg/g) methylene blue uptake in comparison with
Inorganic polymer the other bulk-type geopolymers reported to date (30 h; 15.4 mg/g). The cumulative methylene blue
Geopolymer uptake shown by these innovative spheres (79.7 mg/g) surpasses all other powdered geopolymer ad-
Porosity sorbents, being among the highest values ever reported for geopolymers. The adsorbent was successfully
Dye regenerated and reused eight times. Regeneration was found to negatively affect the MB uptake, but
Waste nevertheless, even after eight regeneration cycles a very high MB removal efficiency (83%) was main-
tained. The use of these bulk-type waste-based geopolymer adsorbents is a low-cost, more eco-friendly,
safer and easier alternative to the use of powdered adsorbents in wastewater treatment systems, since
these ~3 mm spheres may be used directly in packed beds, and were produced using significant amounts
of waste material.
© 2018 Elsevier Ltd. All rights reserved.

1. Introduction its high production cost (Rafatullah et al., 2010) has led to extensive
research for lower cost alternatives. One exciting approach could be
The removal of dyes from industrial wastewaters is of the the use of low cost and eco-friendly inorganic polymers (also
utmost importance, not only because most of them are toxic and known as geopolymers). Inorganic polymers are synthesised at
show carcinogenic properties, but also because treated industrial near-ambient temperatures (Novais et al., 2016b) by chemical
wastewaters may provide a vital source of clean water to mitigate activation of aluminosilicate sources, such as metakaolin (Zhu et al.,
the most pressing concern of our society e water scarcity. In the 2018), fly ash (Novais et al., 2016c) and waste glass (Novais et al.,
next 30 years, around 40% of the world's population will live in 2016a). They have a negatively charged aluminosilicate network,
areas with extreme water scarcity (OECD, 2012). Effective and low balanced by cations such as sodium or potassium, which may in
cost technologies for wastewater treatment are being eagerly pur- turn be exchanged with cations is solution. This feature suggests
sued. Adsorption is considered the most effective, simple and the feasibility of using inorganic polymers as dye or heavy metal
universal technique for water decontamination (Gupta et al., 2012), adsorbent materials. Despite this, the study of the use of geo-
activated carbon being the benchmark adsorbent material (Mestre polymers as adsorbents is fairly recent (Minelli et al., 2018; Novais
et al., 2014). Despite the unique adsorption capacity of this material, et al., 2016d). Moreover, the vast majority of the investigations are
focussed on the use of powdered adsorbents (Falah et al., 2016; Liu
et al., 2016), which cannot be easily recovered, and so cannot be
* Corresponding author. used in field applications or directly in packed beds. Powdered
E-mail address: ruimnovais@ua.pt (R.M. Novais).

https://doi.org/10.1016/j.jclepro.2018.09.265
0959-6526/© 2018 Elsevier Ltd. All rights reserved.
R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362 351

adsorbents may require the use of support materials (e.g. porous Table 1
ceramics, polymer foams) (Zhang et al., 2016) to allow their in- Chemical composition from XRF of metakaolin (MK) and fly ash (FA).

dustrial application or a separation step (e.g. pressure filtration) Oxides (wt.%) MK FA


after wastewater treatment process, both being detrimental to the SiO2 54.40 34.00
wastewater treatment cost, besides increasing the process TiO2 1.55 0.65
complexity. Al2O3 39.40 13.50
Recently, the use of porous geopolymer monoliths for methy- Fe2O3 1.75 4.95
MgO 0.14 3.07
lene blue (MB) extraction from wastewaters was reported by the
CaO 0.10 16.50
authors (Novais et al., 2018a). The MB removal capacity of the MnO 0.01 0.45
monolithic bodies reached 15.4 mg/g, while the adsorbent could be Na2O e 1.52
reused up to 5 times. These promising results demonstrated the K2O 1.03 5.49
SO3 e 2.77
feasibility of using porous bodies, not powders, to extract MB from
P2O5 0.06 1.11
polluted wastewaters. Nevertheless, the adsorbents's removal LOI 2.66 14.30
efficiently dropped significantly, to around 65%, when the MB
initial concentration reached 50 ppm (Novais et al., 2018a). There-
fore, additional investigations addressing the use of bulk porous
Sodium dodecyl sulphate (pore forming agent), polyethylene
geopolymers should shed light on the most influential parameters
glycol (medium used to promote fast curing) and MB (used as
affecting the dye adsorption by the geopolymers. One possibility to
adsorbate) were all supplied by Sigma Aldrich.
enhance the geopolymers' MB adsorption capacity, in comparison
with the use of cylindrical discs (thickness ¼ 3 mm; d ¼ 22 mm)
(Novais et al., 2018a), could be the use of porous geopolymer 2.2. Geopolymer preparation
spheres (GS) (2e3 mm). The authors have recently reported the
synthesis of waste-containing GS by a suspension-solidification The geopolymer slurry was synthesised using an approach
approach (Novais et al., 2017), and their subsequent use as pH described by the authors previously (Novais et al., 2017), in which
regulators in anaerobic digesters (Novais et al., 2018c), while other 15 g of aluminosilicate precursors (50 wt.% metakaolin and 50 wt.%
investigations have studied the use of GS as a heavy metal adsor- FA) were mechanically mixed with 24.38 g of alkaline solution,
bent (Ge et al., 2017) or as a photocatalytic material (Li et al., 2016). 4.15 g of water and 0.75 g of pore forming agent to produce the
Nevertheless, the use of GS as MB adsorbents has never been re- geopolymer slurry. Afterwards, the slurry was injected into a
ported, up to now. This is the first ever investigation regarding the polyethylene glycol medium (~85  C), which promoted the solidi-
use of geopolymer spheres and formulations made from biomass fication of the slurry and the formation of spheres. The FA-based GS
fly ash (FA) waste as MB adsorbent materials from synthetic floated in the polyethylene glycol medium, allowing their easy
wastewaters. The FA waste used was obtained from a local Portu- collection. This suspension-solidification approach has been pre-
guese paper industry, thus enhancing the circular economy aspect viously reported (Novais et al., 2017). After collection, the spheres
of this work. These FA-based GS are expected to show much higher were then washed with distilled water to remove the excess of
surface area than the previously reported geopolymer monoliths sodium hydroxide. Finally the specimens were cured: i) 24 h at
(Novais et al., 2018a) and, therefore, higher MB removal capacity. 40  C and 65% relative humidity in a climatic chamber; and ii) 27
The influence of contact time, MB initial concentration and adsor- days at room temperature, before being used as MB adsorbents. The
bent amount on the removal efficiency of this pollutant by the GS curing procedure was implemented following previous works by
was studied. Desorption and regeneration tests were carried out to the authors (Novais et al., 2017, 2018a).
evaluate the MB extraction efficiency and the feasibility for multi-
ple reuse of the adsorbent. The high removal efficiency shown by 2.3. Methylene blue adsorption tests
these innovative adsorbents, and their multiple recycling, demon-
strates the adsorbents' interesting potential for wastewater treat- Adsorption experiments were carried out to study the influence
ment systems. The present study is a significant step forward in of contact time, MB initial concentration and adsorbent mass on the
comparison with previous investigations (Novais et al., 2018a), removal efficiency of this pollutant by the FA-based GS. Various
clearly demonstrating the performance advantage of the proposed amounts (1.0e2.5 g) of spheres were added to 200 mL of a solution
solution (the use of geopolymer spheres instead of cylindrical discs) containing a specified dye concentration (10e250 ppm), and
for wastewater treatment. magnetically stirred for a predetermined period of time (1e26 h).
Samples were collected from the solution at regular intervals, and
2. Experimental conditions the remaining dye concentration in the liquid evaluated by UV
spectroscopy (Shimadzu UV-3100, JP) by measuring the absorbance
2.1. Materials at a l ¼ 664 nm.
The MB removal by the GS was evaluated by calculating the
Metakaolin (Argical™ M1200S; Univar) and biomass FA waste uptake (qe ) and the removal efficiency percentage (E), respectively,
were used as sources of reactive silica and alumina. The FA was using equations (1) and (2):
sieved, and then only the portion below 63 mm was used in the
geopolymer synthesis. The FA chemical composition, presented in ðC0  Ce Þ
qe ¼ V (1)
Table 1, shows that the waste is a silica-rich material (34 wt.%). m
However, its low alumina content results in a SiO2/Al2O3 ratio ~2.5.
For this reason, metakaolin was also used in the compositions C0  Ce
Eð%Þ ¼  100 (2)
(50 wt.% regarding the aluminosilicate sources) to balance their C0
molar ratios.
For the activation, 100 g of sodium silicate (Chem-Lab, Belgium) where qe quantifies the MB uptake by the GS in mg MB/g geo-
was mixed with 13.22 g of sodium hydroxide (ACS reagent, 97%; polymer, C0 corresponds to the dye initial concentration (mg/L), Ce
Sigma Aldrich), and this mixture was then used as the activator. is the MB equilibrium concentration (mg/L), V the solution volume
352 R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362

and m the adsorbent mass. Eight regeneration cycles were implemented.


Although there are many theories describing the adsorption
equilibrium, those of the Langmuir and Freundlich isotherms are 2.5. Methylene blue desorption tests
the most frequently adopted. The Langmuir isotherm theory fore-
sees monolayer coverage of adsorbate over a homogenous adsor- Desorption tests were performed using specimens after
bent surface. Once the equilibrium is reached, a saturation point is adsorption experiments. For this evaluation, GS which were in
attained and no further adsorption can happen. Sorption is pre- contact with a 50 ppm MB solution were selected. The influence of
sumed to take place at specific homogeneous sites within the their amount (1.0e2.5 g) on the MB desorption capacity was also
adsorbent, this meaning that, once a dye molecule occupies a site, evaluated. After adsorption tests, the spheres were immersed in
no additional adsorption is possible at that specific site (Allen et al., 200 mL of distilled water, and stirred for 24 h (section 3.2.1). Af-
2003): terwards the leached MB concentration was measured by UV
spectroscopy.
KL qm Ce
qe ¼ (3)
1 þ KL Ce 2.6. Materials characterisation and analysis
In equation (3) e a non-linear equation e KL (L/mg) and qm (mg/
The GS microstructure observation before and after MB
g) are the Langmuir isotherm constants. Those can be obtained by
adsorption, and after regeneration, was performed by scanning
making equation (3) linear; this will result in four different types (I,
electron microscopy (SEM) using a Hitachi S4100 system coupled
II, III and IV) of linear equations (Kumar and Sivanesan, 2007). The
with energy dispersive spectroscopy (EDS - Rontec), while the GS
different linearised forms of Langmuir equation are depicted in
morphology was observed using optical analysis (Leica EZ4HD
Table 2.
microscope). ImageJ was used to measure the spheres' diameter
Once the KL value is obtained, the Langmuir isotherm can be
expressed by a separation factor, RL, given by: and length, providing the GS size distribution. For this analysis, 120
spheres were evaluated and the average values presented.
1 The zeta-potential of the FA-based GS was assessed with a
RL ¼ (4) Zetasizer Nano ZS (Malvern) at room temperature.
1 þ KL C0
The aluminosilicate precursor's chemical composition was
Values of 1< RL < 0 represent favourable adsorption (Hajjaji determined by X-ray fluorescence (Philips X'Pert PRO MPD
et al., 2013). spectrometer).
The expression derived by Freundlich (Al Duri and Mckay, 1988), X-ray diffraction (XRD) was used to obtain information about
being an exponential equation [cf equation (5)], assumes that the phase composition of the specimens (i.e. the amorphous and
increasing the adsorbate concentration also increases the concen- the crystalline amounts). For this purpose, full quantitative phase
tration of adsorbate on the adsorbent surface: analyses (FQPA) were assessed using the combined
1
Rietveldereference intensity ratio (RIR) methods, as proposed by
qe ¼ KF C ne (5) Gualtieri (2000) and Gualtieri and Brignoli (2004), using a-Al2O3
(NIST 676a) as an internal standard. This procedure is described in
where KF is the Freundlich constant, and n a parameter which detail in Novais et al. (2018a). XRD data for FQPA were recorded
represents the absence of linearity of the adsorbed quantity in using a q/2q diffractometer (Panalytical Empyrean and X'Pert Pro3,
function of Ce. Equation (5) is normally converted into an alterna- NL), equipped with a linear PIXEL detector (PANalytical), with Cu Ka
tive linear form, thus becoming: radiation(45 kV and 40 mA, 5e80  2q range, with a virtual step scan
of 0.02  2q, and virtual time per step of 200 s). The Rietveld re-
1 finements were accomplished by means of the GSAS software suite,
logqe ¼ logKF þ logCe (6)
n together with its graphical interface EXPGUI (Larson and Von
If 1 < n < 10, then there is favourable adsorption. On the con- Dreele, 2004; Toby, 2001).
trary, larger values of n suggest a stronger interaction between the Fourier-transform infrared spectroscopy (FTeIR) measurements
surface of the adsorbent and adsorbate. When 1/n is equal to 1, a of the adsorbent, before and after MB adsorption, as well as after
linear adsorption has occurred, leading to identical adsorption the regeneration tests, were carried out via a Bruker Tensor 27
energies for all the sites (Febrianto et al., 2009). spectrometer, in attenuated total reflectance (ATR) mode e wave-
number range 4000e350 cm1, 4 cm1 in resolution, 256 scans.

2.4. Adsorbent regeneration tests 3. Results and discussion

The feasibility of reusing the GS as adsorbents after initial MB 3.1. Geopolymer spheres characterisation
adsorption was evaluated only for the specimens showing the
highest MB removal capacity. Samples were heated at 400  C for 2 h Fig. 1a shows a representative optical micrograph of the FA-
to induce MB decomposition. Then the specimens were reused, and based GS, while Fig. 1b presents their size distribution measured
their adsorption ability evaluated using a 50 ppm MB solution. by image analysis. The geopolymers have an elongated spheroidal
shape, with average diameter and length being 2.6 ± 0.2 mm and
2.9 ± 0.3 mm, respectively. These results demonstrate that the
Table 2
production method described here is effective and reproducible,
Linearised forms of the Langmuir equation.
leading to a narrow size distribution. The mm size spheres can be
Linear regression Plot directly used in packed beds and easily retrieved after exhaustion,
Type I Ce/qe versus Ce this being a crucial advantage over powdered adsorbents.
Type II 1/qe versus 1/Ce Fig. S1 (Supplementary Material) shows a high magnification
Type III qe versus qe/Ce SEM micrograph and the corresponding EDS map of the GS surface.
Type IV qe/Ce versus qe
As observed, a geopolymeric gel composed of silica and alumina
R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362 353

Fig. 1. a) Optical characterisation of the FA-based geopolymer spheres and their b) cumulative size distribution (measured using image analysis). Horizontal lines corresponding to
cumulative values of 10, 50 and 90% were included.

was formed, which demonstrates that the alkali activation of the FA and interior pore microstructure. The spheres present a smooth and
and the metakaolin has occurred. This observation is in line with homogeneous surface with several small pores being visible, which
other literature studies which have shown that these two alumi- could allow the MB diffusion into the specimens, thus enhancing
nosilicate sources (FA and metakaolin) can be used as precursors in the adsorption ability of the geopolymers. Fig. 2bec show an
the production of geopolymers (Novais et al., 2016b, 2016c). extremely porous microstructure, in which most of the pores are
Fig. 2 presents SEM micrographs of the spheres' exterior surface closed. The micrographs clearly show the presence of small-sized

Fig. 2. SEM micrographs of the FA-based geopolymer spheres: a) surface and (b, c) interior. Fig. 2d presents the EDS spectra of the geopolymer spheres surface and interior (at two
different positions).
354 R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362

mineral (7.9 wt.%), microcline (14.8 wt.%), traces of anatase


(0.1 wt.%), and 51.1 wt.% amorphous phase. On the other hand,
specimen GS contains a much greater amount of amorphous phase
(82.5 wt.%), whilst the crystalline phases are: aequartz (3.2 wt.%),
calcite (1.7 wt.%), a mica group mineral (5.1 wt.%), microcline
(7.4 wt.%), as well as traces of anatase (0.1 wt.%).

3.2. Methylene blue adsorption tests

3.2.1. Influence of contact time


The evolution with time of MB adsorption by the porous GS was
evaluated using a constant GS amount (1.5 g; number of
spheres ¼ 123 ± 3) and various MB initial concentrations. Consid-
ering the spheroidal shape of the spheres, the average volume per
sphere is roughly 3.27 mm3, so the total volume of spheres used
was ~402 mm3. Fig. 4 shows that in the first 60 min high MB
Fig. 3. Geopolymer spheres zeta potential variation with pH. removal occurs, with removal efficiency ranging from 62 to 77%
depending on the C0 , while afterwards a gentler removal rate is
observed up to 24 h. At this point, the removal efficiency reaches 80
open pores inside the larger-sized closed pores. This is extremely and 93% when the MB initial concentration is 10 and 100 ppm,
relevant, as it will increase the number of active sites available for respectively. Longer sorption times did not induce significant gains,
adsorption. and for that reason 24 h can be considered the equilibrium time.
Fig. 2 also includes the EDS analysis of the spheres' exterior The equilibrium time reported here was compared with other
surface, and two plots corresponding to the spheres' interior. The literature studies performed using powdered and non-powdered
chemical composition changes significantly within the inner part of geopolymers and activated carbons e Fig. 5. It should be high-
the spheres, which is attributed to the distinct chemical composi- lighted that the equilibrium time is affected by several parameters
tion of the aluminosilicate precursors (Table 1), and to the het- such as the adsorbent concentration, the dye initial concentration
erogeneous composition of the FA particles (Novais et al., 2016c). and the adsorption temperature, this hindering the comparison
MK is mainly composed of SiO2 and Al2O3 (account for 93.80 wt.%), between studies which have used various conditions. Anyway, re-
while these two oxides only account for 47.50 wt.% in these FA sults show that the equilibrium time here (24 h) is 20% smaller than
wastes. FA contains several other elements, with CaO (16.50 wt.%), the other bulk geopolymer adsorbent (30 h) reported to date
K2O (5.49 wt.%), Fe2O3 (4.95 wt.%), MgO (3.07 wt.%) and SO3 (Novais et al., 2018a), demonstrating that the use of GS
(2.77 wt.%) being the most abundant.
Fig. 2d also shows a higher sodium content in the spheres'
exterior surface, in comparison with that observed in their inner
part, suggesting that the free sodium remaining in the structure
after geopolymerization migrates/diffuses to the spheres' surface
upon curing. These alkalis will be available for leaching, and are
expected to significantly increase the pH of the MB-containing
wastewater when immersed for long periods (Novais et al.,
2018a). Considering that MB removal efficiency is known to be
affected by the solution pH (Qada et al., 2006), with higher pH
values favouring adsorption, the expected pH increase will further
promote MB uptake by the porous GS.
Fig. 3 presents the zeta potential evolution with pH variation for
the GS. The adsorbent's zeta potential is negative within the studied
pH interval, reaching the highest charge density at pH of ~10.5. The
zeta potential then stabilises until pH values are close to 11.3, before
increasing once more at higher pH values (~12). These results
indicate that pH values between 10.5 and 11.3 will induce the
maximum attraction between the adsorbent and the cationic
adsorbate, and therefore promote higher MB removal efficiency.
Results of FQPA analysis are reported in Table 3. An example of a Fig. 4. Influence of contact time and methylene blue initial concentration on the MB
removal efficiency by the FA-based geopolymer spheres (adsorbent dose: 1.5 g). Dotted
graphic output of a Rietveld refinement is shown in Fig. S2. FA is
lines are a guide for the eye. (For interpretation of the references to colour in this figure
composed of aequartz (11.9 wt.%), calcite (14.1 wt.%), a mica group legend, the reader is referred to the Web version of this article.)

Table 3
Refinement parameters and phase composition of the specimens.

Sample Number of variables Agreement factors Phase composition (wt.%)

R(2F ) (%) Rwp (%) c2 aequartz calcite mica microcline anatase amorphous

FA 38 18.08 4.62 3.19 11.9(2) 14.1(3) 7.9(3) 14.8(5) 0.1(1) 51.1(7)


GS 44 13.16 3.55 2.21 3.2(1) 1.7(1) 5.1(1) 7.4(3) 0.1(1) 82.5(3)
Regenerated GS 44 11.14 3.43 1.86 4.7(1) 4.2(1) 2.6(2) 8.6(5) 0.5(1) 79.4(5)
R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362 355

Fig. 5. Equilibrium time of various methylene blue adsorbents reported in literature. When available the size of the adsorbents was included above the columns (in blue colour) for
comparison (Nasrullah et al., 2018). (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

(d ¼ 2.6 mm) promotes a faster MB uptake by the porous geo-


polymers in comparison with the use of cylindrical discs (thick-
ness ¼ 3 mm; d ¼ 22 mm) (Novais et al., 2018a). This equilibrium
time is even smaller than that reported for some powdered geo-
polymer (kaolin- (Yousef et al., 2009) and FA-based geopolymers (Li
et al., 2006)) and activated carbon (waste rice straw) (Sangon et al.,
2018) adsorbents. Not surprisingly, smaller equilibrium times have
been reported for other powdered adsorbents, geopolymers (Falah
et al., 2016; Khan et al., 2015) and activated carbon (Li et al., 2018).
Nevertheless, these powdered adsorbents cannot be used directly
in packed beds, while the GS can be assembled without the need of
any support material. Although being much larger, and therefore
safer and easier to use in comparison with powdered adsorbents,
the GS still show relatively fast MB sorption, further demonstrating
their potential for wastewater treatment.

3.2.2. Influence of MB initial concentration


Fig. 6. Influence of MB initial concentration on the uptake and removal efficiency of
Fig. 6 presents the uptake and the removal efficiency of the this dye by the porous geopolymer spheres (adsorbent dose: 1.5 g; contact time: 24 h).
porous GS as a function of C0 . The MB uptake sharply increases from
1.1 to 30.1 mg/g when the C0 jumped from 10 to 250 ppm. Inter-
estingly, the removal efficiency increased from 82.3 to 94.3% when diffusion of the dye throughout the samples, the micrographs also
the C0 varied from 10 to 125 ppm, slightly decreasing (to 90.3%) for show that adsorption occurs mainly on the specimen's surface: an
higher C0 values of 200 ppm and over. This behaviour differs from intense bluish hue is observed on the spheres' surface, while a less
our previous investigation (Novais et al., 2018a), in which the intense blue colour is visible in the spheres' interior. This finding
removal efficiency dropped significantly with the increase in C0 . suggests that the MB uptake by the GS could be further enhanced
The trend observed for the removal efficiency can be explained by increasing the porosity of the spheres surface and the connec-
considering the mass transfer resistance between the adsorbate tivity of interior pores. The authors have recently demonstrated the
and the adsorbent, which is affected by the C0 . Higher C0 provides possibility of tuning the superficial and interior porosity of red
an essential driving force to decrease the mass transfer resistance mud-based geopolymer spheres by changing the amount of pore
between the liquid and the solid part (Foo and Hameed, 2012; forming agent (Novais et al., 2018b) or the nature of the binder
Zhang et al., 2011). Optical micrographs of the spheres' micro- (Novais et al., 2018c). Future work will address the optimisation of
structure, shown in Fig. 7, illustrate varying MB diffusion into the the spheres' open porosity to further enhance their adsorption
spheres depending on the C0 , with a more intense blue colour performance.
observed for the GS with up to 175 ppm of MB. Fig. 7a shows that After MB adsorption tests (1.5 g; 24 h) the solutions' pH was
MB did not fully penetrate into the specimens' centre when using a measured to evaluate the impact of the alkalis leaching from the GS
low C0, which hinders the specimen's maximum adsorption level. into the solution. The initial pH of the solution was dependent on
However, increasing the dye concentration decreases the mass the MB initial concentration, ranging from 7.4 to 5.6 when the C0
transfer resistance, which allows a higher penetration into the jumped from 10 to 250 ppm. After the tests, a substantial pH in-
samples and, therefore, higher uptake values. Besides the distinct crease was observed for all compositions, the average value being
356 R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362

Fig. 7. Optical micrographs of the inner part of the geopolymer spheres after MB adsorption illustrating the influence of the MB initial concentration on the MB diffusion throughout
the spheres (real colours are shown). MB initial concentration being: a) 10, b) 25, c) 75, d) 100, e) 175 and f) 200 ppm. (For interpretation of the references to colour in this figure
legend, the reader is referred to the Web version of this article.)

~10. Interestingly, this pH value is extremely close to that found to absent from MK composition. This explains the heterogeneous
induce the highest charge density (10.5) for this innovative adsor- distribution of sulphur (identified in purple in the EDS map) on the
bent (section 3.1), which explains the strong attraction between the spheres' surface before adsorption. The EDS elemental mapping
cationic dye molecules and the negatively charged geopolymer before adsorption also shows sodium agglomerates (identified in
framework. blue), suggesting the presence of unreacted alkalis. After adsorp-
To further understand the MB adsorption by the porous GS, EDS tion, the sodium distribution on the spheres' surface is remarkably
maps from the spheres' surface before and after MB adsorption similar to that of the aluminium and silicon distribution (Fig. 8b).
tests were collected, and are shown in Fig. 8. According to the XRF This observation suggests that the free sodium observed in the GS
data, the FA contains around 3 wt.% of SO3, while this element is before adsorption has leached to the solution, while the remaining

Fig. 8. EDS elemental mapping of FA-based geopolymer spheres' surface a) before and b) after methylene blue adsorption tests. (For interpretation of the references to colour in this
figure legend, the reader is referred to the Web version of this article.)
R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362 357

sodium in acting as a charge-balancing cation for the negatively (1355 and 1335 cm1, and ~1410 cm1, respectively) (Coates, 2006)
charged aluminosilicate network, which explains the above- occurred, providing clear evidence of MB adsorption.
mentioned pH increase. After MB adsorption, the sulphur distri-
bution also changes, with two clear features being distinguished: i) 3.2.3. Influence of adsorbent concentration
a homogeneous distribution associated with the geopolymer The effect of the adsorbent amount on the MB uptake and
structure (derived from the FA composition); and ii) the presence of removal efficiently by the GS was evaluated to determine the op-
small sulphur concentrations (not observed before) which were timum adsorbent dosage, and results are shown in Fig. 10. The MB
associated with the MB adsorption by the porous GS (MB molecular uptake and the removal efficiency were affected differently by the
formula: C16H18ClN3S). increase in the adsorbents amount, the uptake drops and the
FT-IR spectra are depicted in Fig. 9. They all show similar char- removal efficiency increases (exceptions to this general trend will
acteristics: the strong asymmetric band, having its maximum at be discussed below) when rising the GS amount. Increasing the
around 3350 cm1, is due to adsorbed molecular H2O, which also adsorbents amount increases the number of available adsorption
generates the feature at around 1630 cm1, assigned to HeOeH sites, therefore increasing the amount of adsorbed MB (i.e. removal
bending (Tobaldi et al., 2010). The band located at ~980 cm1 be- efficiency). However, the amount adsorbed per unit of mass (i.e.
longs to the asymmetric AleOeAl/SieOeSi stretching (Kumar et al., uptake) decreases which is explained by the presence of unsatu-
2015). The band at ~1440 cm1 can be assigned to sodium car- rated adsorption sites (Aljeboree et al., 2017). The decrease in the
bonate, due to the reaction of residual sodium with atmospheric MB uptake when rising the adsorbents amount has been previously
CO2 (Cheng-Yong et al., 2017). reported (Liu et al., 2016).
After being in contact with MB (24 h), additional features Results also showed that the MB removal efficiency was affected
appeared (cf the blue continuous line in Fig. 9). The band centred at by the dye initial concentration: i) C0 ¼ 10 ppm e the removal ef-
~1610 cm1, belonging to the stretching vibration of aromatic rings, ficiency significantly dropped from 82 to 75% when the spheres
and features due to methyl and methylene bending vibrations content was above 1.5 g; ii) C0 ¼ 50 ppm e the removal efficiency
increased up to 2.0 g, and then slightly decreased (~1%) afterwards;
iii) C0 ¼ 100 ppm e the removal efficiency increased steadily with
the rise of the adsorbent amount, from ~88 up to ~96%, until a
plateau was reached. These different trends are better illustrated in
Fig. 11 where the removal efficiency versus contact time is shown.
As depicted, the sorption kinetics are MB concentration-dependent,
with higher C0 inducing a faster MB uptake, this being independent
from the amount of adsorbents used. These differences were
associated with the above-mentioned mass transfer resistance. In
any case, for the higher C0 values, an increase in the amount of
spheres from 1.0 to 1.5 g enhanced the removal efficiency by around
5%. Increasing the quantity of spheres used to 2.0 and 2.5 g of
spheres promoted even higher removal efficiencies, an increase of
~2% and ~3%, respectively, in comparison with the use of 1.5 g.
Despite the slight gain in performance when using greater quan-
tities of spheres, the use of 1.5 g of spheres seems to be the most
cost-effective solution (E ¼ 93%). This is the reason why 1.5 g of
spheres was used as a proxy for the isotherm model studies, see
below.

Fig. 9. FTIR spectra of the geopolymer spheres before, after MB adsorption, (contact
3.3. Isotherm model studies
time ¼ 24 h; adsorbent dose: 1.5 g; C0 ¼ 200 ppm), as well as being regenerated. Note:
the noise in the 2450e2000 cm1 region is due to atmospheric CO2. Linear regression results, for both the Freundlich and Langmuir

Fig. 10. Influence of the adsorbent amount on a) the MB uptake and b) removal efficiency by the FA-based geopolymer spheres for various concentrations of MB (contact time: 24 h).
358 R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362

Fig. 11. Influence of the adsorbent amount and MB initial concentration on the removal efficiency kinetics.

isotherm equations, are reported in Fig. S3. As seen in Fig. S3, the Table 4
linearisation resulted in a two-step trend, with the turning point Isotherm constants for MB adsorption onto GS (1.5 g of adsorbent in 0.2 L MB
solution).
(for the system considered here, i.e., 1.5 g of GS as the adsorbent
medium, MB as the solute, in 0.2 L of solution) being located at Type I Type II Type III Type IV
~100 ppm of MB (Freundlich model) and 125 ppm of MB (Langmuir Langmuir constants, MB concentration range: 10e125 ppm
model).
qm (mg/g) 4.028 3.619 4.070 4.131
This is better visualised in Table 4, where the Langmuir (Type I to KL (L/mg) 0.110 0.116 0.110 0.109
IV) and Freundlich constants are listed, considering MB concen- R2 0.954 0.991 0.994 0.994
tration intervals of 10e100 and 125e250 ppm for the Freundlich Freundlich constants, MB concentration range: 10e100 ppm
isotherm model, and 10e125 and 125e250 ppm for the Langmuir
KF (mg/g)(L/g)n 0.276
model. As derived from Table 4, at relatively low MB concentrations 1/n 1.895
(10e125 ppm), the data follows the Freundlich isotherm equation. R2 0.993
The determination coefficient of the Freundlich model, at this
Langmuir constants, MB concentration range:125e250 ppm
concentration range, is 0.993; although the R2 of the Langmuir
qm (mg/g) 45.434 45.809 43.562 46.960
models (Types III and IV) are slightly better than that (i.e. 0.994), the
KL (L/mg) 0.076 0.074 0.084 0.072
Langmuir isotherm constants are all negative (for all of the four R2 0.959 0.996 0.822 0.822
type of linear methods) e this being physically meaningless. On the
Freundlich constants, MB concentration range: 125e250 ppm
contrary, when dealing with higher MB concentrations
(i.e.  100 ppm), the Langmuir isotherm model became more KF (mg/g)(L/g)n 6.227
1/n 0.495
applicable; in particular the Type II has a very good R2 (0.996), with
R2 0.946
an adsorption capacity qm equal to 45.809 mg/g, this suggesting the
prevalence of monolayer adsorption as the dominant sorption The bold identifies the isotherm model that better fits the experimental results.

mechanism on that macroporous adsorbent (Rouquerol et al.,


1998). This is to be expected, as transforming a non-linear equa-
the experimental data in a better way at relatively low concentra-
tion into a linear form implies altering the error structure and the
tions (i.e.  100 ppm), whilst those for the Langmuir had the ten-
error variance (Allen et al., 2003). This might explain the observa-
dency to fit better at concentrations 125 ppm (Richter et al., 1989).
tions that linear Freundlich parameters gave isotherms which fitted
R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362 359

These results are in good agreement with the experimental data


discussed in section 3.2.3. As clearly shown by the optical micro-
graphs (Fig. 7), at low concentration a heterogeneous MB adsorp-
tion is attained, consistent with the better fitting achieved with the
Freundlich model; while at high C0 the MB adsorption is more
homogeneous, in particular at the spheres' surface, this explaining
the better fitting of the Langmuir model in this case. The differences
in adsorption depending on the MB initial concentration were
attributed to the decrease in the mass transfer resistance at high C0 ,
as discussed in section 3.2.2.

3.4. Geopolymer spheres recyclability

The feasibility of reusing the GS in multiple adsorption cycles


was evaluated by using the spheres (1.5 g) that showed the highest
MB uptake (30.1 mg/g) and results for the MB uptake by the GS are
shown in Fig. 12. In the first adsorption cycle C0 was 250 ppm.
However, after the 1st regeneration cycle, the GS removal ability for
such a high MB initial concentration was limited. For this reason, Fig. 13. Influence of the number of regeneration cycles on the MB removal efficiency
by the geopolymer spheres. Note: the data in the first hours for the “1st adsorption
after 24 h adsorption the MB equilibrium concentration was still
(200 ppm)” was above the UVeVis spectrometer detection limit.
above the UV detection limit. Considering this performance, the
following regeneration cycles were performed using a lower C0 of
50 ppm. samples (non- and regenerated) was similar, suggesting that
There are three possible explanations for this drop in the use of higher C0 for the regeneration tests could be
performance: possible, but that the equilibrium time would surpass 24 h.
ii) The lower pH attained after adsorption/regeneration cycles.
i) Modification of the adsorption kinetics due to the greater As previously mentioned, the solution pH after the 1st
number of spheres used in the subsequent recycling tests. adsorption test was ~10. However, the solutions' pH signifi-
After the 1st adsorption test (24 h immersion in MB solution) cantly dropped after subsequent regeneration cycles, from
the mass of the spheres was significantly reduced by ~33%, pH ¼ 8.9 after the 1st, and down to pH ¼ 7.0 after the 8th
which was associated with the loss of the alkalis released regeneration cycle. The reason for the pH drop when the
from the GS. Additionally, the 1st thermal treatment (section number of regeneration cycles increases is related to the
2.5) further reduced the adsorbents' weight by ~16%, asso- amount of alkalis available for leaching. As discussed above
ciated to the extraction of physically adsorbed water, (see section 3.1), the geopolymers contain within their
meaning that after the first adsorption/regeneration cycle structure free alkalis which are available for leaching. How-
the spheres had lost ~49% of their mass. Thus, to use the same ever, as the number of regeneration cycles increases, the al-
amount of adsorbent (1.5 g), a significantly higher number of kalis availability decreases, since most of it has already
spheres had to be used (~241 spheres instead of 123), which leached into the solution, thus explaining the observed pH
altered the adsorption kinetics. Fig. 13 shows a lower sorp- drop. This lower pH reduces the attraction between the
tion rate in the first couple of hours for the regenerated cationic dye and the geopolymers' negatively charged
samples (1.5 g; 241 spheres) in comparison with the speci- aluminosilicate network, thus negatively affecting the MB
mens that were solely immersed in a 50 ppm MB solution uptake.
(without thermal regeneration; 1.5 g; 123 spheres). Never- iii) Incomplete degradation of MB during the thermal regener-
theless, after 24 h the removal efficiency between the ation. As demonstrated above (Fig. 7), MB adsorption occurs
mainly at the spheres' surface, and to a lower degree in their
inner part (when MB diffuses into the pores). During thermal
regeneration, the MB adsorbed on the spheres' surface is
removed; however, the MB adsorbed inside the pores might
not be completely removed, which would explain the per-
formance drop. The FTIR spectrum of the regenerated sample
(milled before the measurement) does not shown the pres-
ence of MB (Fig. 9), suggesting that the thermal treatment
was effective. Nevertheless, it's possible that the remaining
MB moieties in the sample are below the equipment detec-
tion limit, and for that reason the possibility of incomplete
MB degradation cannot be ruled out. Future work will
consider extending the thermal treatment to promote com-
plete MB degradation. Furthermore, FQPA analysis of the
regenerated specimen showed a decrease in the amorphous
phase amount (from 82.5 to 79.4 wt.%, cf Table 3), likely due
to the thermal treatment. This also favoured the on-going
process of carbonatation, the calcite content increased from
1.7 wt.% in GS to 4.2 wt.% in the regenerated specimen.
Fig. 12. Influence of the number of regeneration cycles on the MB uptake by the FA-
based geopolymer spheres. The results shown in Fig. 12 clearly demonstrate the feasibility of
360 R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362

Fig. 14. SEM micrographs of the spheres' surface (a) and inner part (b) after eight regeneration cycles illustrating the microstructural changes induced in the geopolymer spheres by
the multiple adsorption and thermal regeneration cycles. Fig. 14c presents the EDS spectrum of the geopolymer sphere surface.

reusing this innovative adsorbent in multiple MB adsorption cycles, that after 9 cycles (1 adsorption þ 8 adsorptions after regeneration)
this being a crucial advantage over benchmark activated carbons, the sample losses ~61% of its initial mass (49% þ 12%). Nevertheless,
whose recovery after use is extremely challenging. After the 1st these results suggest that additional regeneration cycles could be
regeneration cycle, seven other cycles (adsorption and thermal performed. This will be considered in future work.
regeneration) were performed without any further performance The cumulative MB uptake of the GS (79.7 mg/g), shown in
compromise. At the end of the eight regeneration cycles, the GS Fig. 12, demonstrates the interesting potential of this bulk adsor-
maintained their integrity, as demonstrated by the SEM micro- bent for wastewater treatment systems. The removal efficiency
graphs shown in Fig. 14, although an increase in porosity was throughout the nine cycles was always above 83% (results not
observed. A continuous and gradual mass reduction (reaching 12% shown by the sake of brevity), much higher than the other study
after the 8th cycle) was observed throughout the tests. This means addressing the use of bulk-type geopolymer adsorbent whose

Table 5
Maximum methylene blue adsorption capacity of various adsorbents reported in literature.

Adsorbent shape Material Removal efficiency (%) qe (mg/g) Reference


a
powder Coal fly ash geopolymer 90 0.7 (Zhang and Liu, 2013)
Phosphoric acid MK-based geopolymer 100 a 3.0 (Khan et al., 2015)
Coal fly ash Not given 3.8 (Wang et al., 2015)
Acid treated zeolite 95.3 4.0e4.5 (Hor et al., 2016)
Acid treated coal fly ash Not given 8.0 (Wang et al., 2015)
Acrylic fibrous waste activated carbon Not given 8.8 (Naeem et al., 2017)
Cu2O-geopolymer Not given 14.8 (Falah et al., 2015)
Cu2O/TiO2 geopolymer Not given 14.8 (Falah and Mackenzie, 2015)
Cu2O/TiO2-CTAB geopolymer Not given 19.7 (Falah et al., 2016)
Kaolin geopolymer 90 a 25.6 (Yousef et al., 2009)
Zeolite Not given 33.5 (Rida et al., 2013)
Activated lignin-chitosan 88 36.3 (Albadarin et al., 2017)
Coal fly ash geopolymer 80 a 38.4 (Li et al., 2006)
Ficus carica bast activated carbon Not given 48 (Pathania et al., 2017)
Coal fly ash geopolymer 25 a 50.7 (Liu et al., 2016)
Fe3O4 nanoparticles Not given 91.9 (Ghaedi et al., 2015)
Palm shell activated carbon Not given 163 (Wong et al., 2016)
Waste rice straw activated carbon Not given 528 (Sangon et al., 2018)
Bulk-type Biomass FA-geopolymer monoliths 64.8 15.4 (1st adsorption) (Novais et al., 2018a)
48.4 109.0 (6th adsorption)
FA-based geopolymer spheres 90.3 30.1 (1st adsorption) This work
84.9 79.7 (9th adsorption)

The font in bold identifies the results obtained in this work.


a
Maximum removal efficiency values extrapolated from the experimental data.
R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362 361

removal efficiency after the 1st adsorption test was only 55%, The MB uptake by the porous geopolymer spheres is affected by the
dropping to values below 50% when using the regenerated samples dye initial concentration, adsorbent amount and contact time. Re-
(up to 5 cycles) (Novais et al., 2018a). sults showed a faster (20% smaller equilibrium time) and higher MB
The maximum (experimental) adsorption capacity observed for uptake (twofold increase) in comparison with the other bulk-type
the GS (79.7 mg/g) was compared with that reported for other MB geopolymer reported to date, demonstrating the performance
adsorbents, and results are shown in Table 5. The GS show very high advantage of using spheres (d ¼ 2.6 mm) instead of cylindrical discs
MB removal capacity, surpassing all reported values for powdered (thickness ¼ 3 mm; d ¼ 22 mm) for wastewater decontamination.
geopolymers, such as those prepared from coal FA (Li et al., 2006; After adsorption tests, the geopolymer spheres could be re-
Liu et al., 2016), MK (Khan et al., 2015), Cu2O-modified geopolymers generated and reused (up to 8 cycles) resulting in a very high cu-
(Falah et al., 2016; Falah and Mackenzie, 2015), kaolin-based mulative MB uptake (79.7 mg/g) which surpasses all other
(Yousef et al., 2009), and acid treated zeolites (Hor et al., 2016), powdered geopolymer adsorbents. These remarkable materials
among others (Rida et al., 2013; Wang et al., 2015). The GS may be used directly in wastewater systems (e.g. packed beds), and
adsorption capacity is inferior to that previously reported by the are easily retrieved after exhaustion, these being crucial advantages
authors when using geopolymer monoliths (Novais et al., 2018a). over conventional powdered adsorbents. The proposed solution
However, in this previous investigation the removal efficiency was (use of mm size geopolymer spheres instead of powdered adsor-
48.4% and the sorption time was 30 h (Novais et al., 2018a), while bents) increases the simplicity of wastewater treatment systems,
here a significantly higher efficiency (84.9%) with lower sorption being a safer alternative to conventional powdered adsorbents,
time (24 h) was achieved. Moreover, the GS are synthesised whose recovery after use is extremely challenging.
through one-step, while two steps are required for the monoliths Moreover, this innovative adsorbent also promotes waste min-
production (i) synthesis of the porous bodies; and ii) cutting of imisation, since biomass FA wastes were used to partially replace
cylindrical discs; thus, a simplified strategy is used here. (50 wt.%) commercial metakaolin in the geopolymer spheres' syn-
Table 5 also shows that the GS adsorption capacity is superior to thesis, which is aligned with the need to reduce the consumption of
that reported for Ficus carica bast (Pathania et al., 2017) and acrylic virgin raw materials and prevent landfill disposal of wastes.
fibrous waste (Naeem et al., 2017) activated carbons, whilst being
inferior to that of Fe3O4 nanoparticles (Ghaedi et al., 2015), and to Acknowledgements
palm shell (Wong et al., 2016) and waste rice straw (Sangon et al.,
2018) activated carbons. Nevertheless, the proposed strategy (use This work was developed within the scope of the project
of mm size spheres instead of powders) may allow the adsorbent's CICECO-Aveiro Institute of Materials, POCI-01-0145-FEDER-007679
direct use in wastewater treatment facilities, and its easy recovery (FCT Ref. UID/CTM/50011/2013), financed by national funds
after exhaustion, while the recovery of nanoparticles or activated through the FCT/MEC and when appropriate co-financed by FEDER
carbons after adsorption is a complex and expensive procedure. under the PT2020 Partnership Agreement. R.C. Pullar thanks the
Furthermore, the GS were synthesised are near room temperature FCT for funding under grant IF/00681/2015.
(24 h at 40  C), preventing the need for very high processing tem-
peratures (700  C) typically employed in the production of acti-
Appendix A. Supplementary data
vated carbons (Sangon et al., 2018), and thus being an
environmental friendlier approach.
Supplementary data to this article can be found online at
The outstanding recyclability of the GS, assuring very high cu-
https://doi.org/10.1016/j.jclepro.2018.09.265.
mulative MB adsorption capacity, associated with the possibility of
being directly used in wastewater systems without the need for
support materials, makes this bulk-adsorbent an excellent and safer References
alternative to conventional powdered adsorbents for wastewater
Albadarin, A.B., Collins, M.N., Naushad, M., Shirazian, S., Walker, G., Mangwandi, C.,
decontamination. Future work will evaluate the possibility of 2017. Activated lignin-chitosan extruded blends for efficient adsorption of
reusing the GS as aggregates in the production of eco-friendly methylene blue. Chem. Eng. J. 307, 264e272.
Al Duri, B., Mckay, G., 1988. Basic dye adsorption on carbon using a solid-phase
mortars, after their exhaustion as adsorbent material, which
diffusion model. Chem. Eng. J. 38, 23e31.
would contribute towards the circular economy, besides decreasing Aljeboree, A.M., Alshirifi, A.N., Alkaim, A.F., 2017. Kinetic and equilibrium study for
the carbon footprint of the spheres. the adsorption of textile dyes on coconut shell activated carbon. Arab. J. Chem.
10, S3381eS3393.
Allen, S.J., Gan, Q., Matthews, R., Johnson, P.A., 2003. Comparison of optimized
3.5. Methylene blue desorption tests isotherm models for basic dye adsorption by kudzu. Bioresour. Technol. 88,
143e152.
To evaluate the MB fixation onto the GS desorption tests were Cheng-Yong, H., Yun-Ming, L., Al Abdullah, M.M., Hussin, K., 2017. Thermal resis-
tance variations of fly ash geopolymers: foaming responses. Sci. Rep. 7, 45355.
performed in distilled water. All specimens showed low leaching Coates, J., 2006. Interpretation of infrared spectra, a practical approach. In:
values, ranging from 0.21 to 0.57%, suggesting a strong fixation of Meyers, R.A. (Ed.), Encyclopedia of Analytical Chemistry. John Wiley & Sons Ltd,
the dye onto the spheres. The quantity of spheres used did not Chichester, pp. 10815e10837.
El Qada, E.N., Allen, S.J., Walker, G.M., 2006. Adsorption of methylene blue onto
significantly affect the MB desorption level, although a slight activated carbon produced from steam activated bituminous coal: a study of
decrease was observed when increasing the amount of spheres: equilibrium adsorption isotherm. Chem. Eng. J. 124, 103e110.
0.57% (1.0 g), 0.37% (1.5 g), 0.22% (2.0 g) and 0.21% (2.5 g). These Falah, M., Mackenzie, K.J.D., Hanna, J.V., Page, S.J., 2015. Novel photoactive inorganic
polymer composites of inorganic polymers with copper (I) oxide nanoparticles.
results indicate that leaching in water medium is ineffective, J. Mater. Sci. 50, 7374e7383.
meaning that the adsorbent can be used for longer equilibrium Falah, M., Mackenzie, K.J.D., Knibbe, R., Page, S.J., Hanna, J.V., 2016. New composites
times than those employed here (Section 3.2.1) to ensure the of nanoparticle Cu (I) oxide and titania in a novel inorganic polymer (geo-
polymer) matrix for destruction of dyes and hazardous organic pollutants.
maximum MB removal (saturation of the spheres).
J. Hazard Mater. 318, 772e782.
Falah, M., Mackenzie, K.J.D., 2015. Synthesis and properties of novel photoactive
4. Conclusions composites of P25 titanium dioxide and copper (I) oxide inorganic polymers.
Ceram. Int. 41, 13702e13708.
Febrianto, J., Kosasih, A.N., Sunarso, J., Ju, Y.-H., Indraswati, N., Ismadji, S., 2009.
In this investigation, and for the first time, FA-based geopolymer Equilibrium and kinetic studies in adsorption of heavy metals using biosorbent:
spheres were evaluated as a methylene blue adsorbent material. a summary of recent studies. J. Hazard Mater. 162, 616e645.
362 R.M. Novais et al. / Journal of Cleaner Production 207 (2019) 350e362

Foo, K.Y., Hameed, B.H., 2012. Dynamic adsorption behaviour of methylene blue Porous biomass fly ash-based geopolymers with tailored thermal conductivity.
onto oil palm shell granular activated carbon prepared by microwave heating. J. Clean. Prod. 119, 99e107.
Chem. Eng. J. 203, 81e87. Novais, R.M., Buruberri, L.H., Seabra, M.P., Bajare, D., Labrincha, J.A., 2016c. Novel
Ge, Y., Cui, X., Liao, C., Li, Z., 2017. Facile fabrication of green geopolymer/alginate porous fly ash-containing geopolymers for pH buffering applications. J. Clean.
hybrid spheres for efficient removal of Cu (II) in water: batch and column Prod. 124, 345e404.
studies. Chem. Eng. J. 311, 126e134. Novais, R.M., Buruberri, L.H., Seabra, M.P., Labrincha, J.A., 2016d. Novel porous fly
Ghaedi, M., Hajjati, S., Mahmudi, Z., Tyagi, I., Agarwal, S., Maity, A., Gupta, V.K., 2015. ash-containing geopolymer monoliths for lead adsorption from wastewaters.
Modeling of competitive ultrasonic assisted removal of the dyes e methylene J. Hazard Mater. 318, 631e640.
blue and safranin-O using Fe3O4 nanoparticles. Chem. Eng. J. 268, 28e37. Novais, R.M., Carvalheiras, J., Seabra, M.P., Pullar, R.C., 2018b. J.A. Labrincha Inno-
Gualtieri, A.F., 2000. Accuracy of XRPD using combined Rietveld-RIR method. vative application for bauxite residue: red mud-based inorganic polymer
J. Appl. Crystallogr. 33, 267e278. spheres as pH regulators. J. Hazard Mater. 358C, 69e81.
Gualtieri, A.F., Brignoli, G., 2004. Rapid and accurate quantitative phase analyses Novais, R.M., Gameiro, T., Carvalheiras, J., Seabra, M.P., Tarelho, L.A.C., Labrincha, J.A.,
using a fast detector. J. Appl. Crystallogr. 37, 8e13. Capela, I., 2018c. High pH buffer capacity biomass fly ash-based geopolymer
Gupta, V.K., Ali, I., Saleh, T.A., Nayak, A., Agarwal, S., 2012. Chemical treatment spheres to boost methane yield in anaerobic digestion. J. Clean. Prod. 178,
technologies for waste-water recycling e an overview. RSC Adv. 2, 6380e6388. 258e267.
Hajjaji, W., Ganiyu, S.O., Tobaldi, D.M., Andrejkovi cova, S., Pullar, R.C., Rocha, F., Novais, R.M., Seabra, M.P., Labrincha, J.A., 2017. Geopolymer spheres as novel pH
Labrincha, J.A., 2013. Natural Portuguese clayey materials and derived TiO2- buffering materials. J. Clean. Prod. 143, 1114e1122.
containing composites used for decolouring methylene blue (MB) and orange II OECD (Organisation for Economic Co-operation and Development), 2012. 2012
(OII) solutions. Appl. Clay Sci. 83e84, 91e98. OECD (Organisation for Economic Co-operation and Development) OECD
Hor, K.Y., Chee, J.M.C., Chong, M.N., Jin, B., Saint, C., Poh, P.E., Aryal, R., 2016. Eval- Environmental Outlook to 2050: the Consequences of Inaction. OECD Publish-
uation of physicochemical methods in enhancing the adsorption performance ing, Paris. https://doi.org/10.1787/9789264122246-en.
of natural zeolite as low-cost adsorbent of methylene blue from wastewater. Pathania, D., Sharma, S., Singh, P., 2017. Removal of methylene blue by adsorption
J. Clean. Prod. 118, 197e209. onto activated carbon developed from Ficus carica bast. Arab. J. Chem. 10,
Khan, M.I., Min, T.K., Azizli, K., Sufian, S., Man, Z., Ullah, H., 2015. Effective removal of S1445eS1451. https://doi.org/10.1016/j.arabjc.2013.04.021.
methylene blue from water using phosphoric acid based geopolymers. RSC Adv. Rafatullah, M., Sulaiman, O., Hashim, R., Ahmad, A., 2010. Adsorption of methylene
5, 61410e61420. blue on low-cost adsorbents: a review. J. Hazard Mater. 177, 70e80.
Kumar, K.V., Sivanesan, S., 2007. Isotherms for malachite green onto rubber wood Richter, E., Schütz, W., Myers, A.L., 1989. Effect of adsorption equation on prediction
(hevea brasiliensis) sawdust: comparison of linear and non-linear methods. of multicomponent adsorption equilibria by the ideal adsorbed solution theory.
Dyes Pigments 72, 124e129. Chem. Eng. Sci. 44, 1609e1616.
Kumar, S., Krista ly, F., Mucsi, G., 2015. Geopolymerisation behaviour of size frac- Rida, K., Bouraoui, S., Hadnine, S., 2013. Adsorption of methylene blue from aqueous
tioned fly ash. Adv. Powder Technol. 26, 24e30. solution by kaolin and zeolite. Appl. Clay Sci. 83e84, 99e105.
Larson, A.C., Von Dreele, R.B., 2004. General Structure Analysis System (GSAS). Rouquerol, F., Rouquerol, J., Sing, K.S.W., 1998. Adsorption by Powders and Porous
National Laboratory Report LAUR, Los Alamos, pp. 86e748. Solids, Principles, Methodology and Applications, first ed. Academic Press, New
Li, C.-m., He, Y., Tang, Q., Wang, K.-t., Cui, X.-m., 2016. Study of the preparation of York. 978-0-08-052601-0.
CdS on the surface of geopolymer spheres and photocatalyst performance Sangon, S., Hunt, A.J., Attard, T.M., Mengchang, P., Ngernyen, Y., Supanchaiyamat, N.,
Materials. Chem. Phys. 178, 204e2010. 2018. Valorisation of waste rice straw for the production of highly effective
Li, L., Wang, S., Zhu, Z., 2006. Geopolymeric adsorbents from fly ash for dye removal carbon based adsorbents for dyes removal. J. Clean. Prod. 172, 1128e1139.
from aqueous solutions. J. Colloid Interface Sci. 300, 52e59. 
Tobaldi, D.M., Tucci, A., Skapin, A.S., Esposito, L., 2010. Effects of SiO2 addition on
Liu, Y., Yan, C., Zhang, Z., Gong, Y., Wang, H., Qiu, X., 2016. A facile method for TiO2 crystal structure and photocatalytic activity. J. Eur. Ceram. Soc. 30,
preparation of floatable and permeable fly ash-based geopolymer block. Mater. 2481e2490.
Lett. 185, 370e373. Toby, B.H., 2001. EXPGUI, a graphical user interface for. GSAS J. Appl. Crystallogr. 34,
Li, Z., Wang, G., Zhai, K., He, C., Li, Q., Guo, P., 2018. Methylene blue adsorption from 210e213.
aqueous solution by loofah sponge-based porous carbons. Colloids Surf., A 538, Wang, S., Boyjoo, Y., Choueib, A.A., 2015. A comparative study of dye removal using
28e35. fly ash treated by different methods. Chemosphere 60, 1401e1407.
Mestre, A.S., Pires, R.A., Aroso, I., Fernandes, E.M., Pinto, M.L., Reis, R.L., Wong, K.T., Eu, N.C., Ibrahim, S., Kim, H., Yoon, Y., Jang, M., 2016. Recyclable
Andrade, M.A., Pires, J., Silva, S.P., Carvalho, A.P., 2014. Activated carbons pre- magnetite-loaded palm shell-waste based activated carbon for the effective
pared from industrial pre-treated cork: sustainable adsorbents for pharma- removal of methylene blue from aqueous solution. J. Clean. Prod. 115, 337e342.
ceutical compounds removal. Chem. Eng. J. 253, 408e417. https://doi.org/10.1016/j.jclepro.2015.12.063.
Minelli, M., Papa, E., Medri, V., Miccio, F., Benito, P., Doghieri, F., Landi, E., 2018. Yousef, R.I., El-Eswed, B., Alshaaer, M., Khalili, F., Khoury, H., 2009. The influence of
Characterization of novel geopolymer - zeolite composites as solid adsorbents using Jordanian natural zeolite on the adsorption, physical, and mechanical
for CO2 capture. Chem. Eng. J. 341, 505e515. properties of geopolymers products. J. Hazard Mater. 165, 379e387.
Naeem, S., Baheti, V., Wiener, J., Marek, J., 2017. Removal of methylene blue from Zhang, W., Dong, L., Yan, H., Li, H., Jiang, Z., Kan, X., Yang, H., Li, A., Cheng, R., 2011.
aqueous media using activated carbon web. J. Textil. Inst. 108, 803e811. Removal of methylene blue from aqueous solutions by straw based adsorbent in
Nasrullah, A., Bhat, A.H., Naeem, A., Isa, M.H., Danish, M., 2018. High surface area a fixed-bed column. Chem. Eng. J. 173, 429e436.
mesoporous activated carbon-alginate beds for efficient removal of methylene Zhang, Y., Liu, L., 2013. Fly ash-based geopolymer as a novel photocatalyst for
blue. Int. J. Biol. Macromol. 107, 1792e1799. degradation of dye from wastewater. Particuology 11, 353e358.
Novais, R.M., Ascens~ ao, G., Seabra, M.P., Labrincha, J.A., 2016a. Waste glass from end- Zhang, Z., Li, L., He, D., Ma, X., Yan, C., Wang, H., 2016. Novel self-supporting zeolitic
of-life fluorescent lamps as raw materials in geopolymers. Waste Manag. 52, block with tunable porosity and crystallinity for water treatment. Mater. Lett.
245e255. 178, 151e154.
Novais, R.M., Ascensa ~o, G., Tobaldi, D.M., Seabra, M.P., Labrincha, J.A., 2018a. Zhu, W., Rao, X.H., Liu, Y., Yang, E.-H., 2018. Lightweight aerated metakaolin-based
Biomass fly ash geopolymer monoliths for effective methylene blue removal geopolymer incorporating municipal solid waste incineration bottom ash as
from wastewaters. J. Clean. Prod. 171, 783e794. gas-forming agent. J. Clean. Prod. 177, 775e781.
Novais, R.M., Buruberri, L.H., Ascensa ~o, G., Seabra, M.P., Labrincha, J.A., 2016b.

You might also like