You are on page 1of 10

Journal of Cleaner Production 148 (2017) 905e914

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

A comparative study of sodium/hydrogen titanate nanotubes/


nanoribbons on destruction of recalcitrant compounds and
sedimentation
Kunlanan Kiatkittipong a, *, Suttichai Assabumrungrat b
a
Department of Chemical Engineering, Faculty of Engineering, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand
b
Center of Excellence in Catalysis and Catalytic Reaction Engineering, Department of Chemical Engineering, Faculty of Engineering, Chulalongkorn
University, Bangkok 10330, Thailand

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

Article history: The performances of sodium titanate nanotubes (Na-TT), hydrogen titanate nanotubes (H-TT), sodium
Received 6 August 2016 titanate nanoribbons (Na-TR) and hydrogen titanate nanoribbons (H-TR) on adsorption, photo-
Received in revised form degradation of methylene blue (MB) and sedimentability were investigated. The intrinsic property of Na-
6 February 2017
TT with high hydroxyl groups (anionic) provided a strong electrostatic interaction between the MB
Accepted 6 February 2017
Available online 6 February 2017
(cationic) under dark condition. The equilibrium and kinetic adsorption were fitted with Langmuir
isotherm models and the pseudo-second-order model, respectively. That is, Na-TT showed the best
maximum adsorption capacity of 178.6 mg/g and adsorption kinetics of 0.119 g/mg$min. When the light
Keywords:
Titanate
is turned on, H-TT showed advance its photocatalytic reaction rate of 1.16  102 min1 compared with
Nanotube Na-TT (the reaction rate of 1.02  102 min1). This attributed the saturated MB adsorption on Na-TT
Nanoribbon obstructed capacity to absorb photons, and thus the reaction gradually occurred. For the separation of
Adsorption photocatalysts from the suspension by adding alum as coagulant, Na-TT afforded the best sedimentability
Photocatalytic activity of which remaining Na-TT in water was approximately 5%. This study demonstrates that adsorption,
Sedimentation degradation and sedimentability were governed by the surface area and crystal structure, which revealed
increasing the performance in the order TiO2 < H-TR < Na-TR < H-TT < Na-TT.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction low concentrations, causing human health and environment


problems. Therefore, the effective removal of bio-recalcitrant
Wastewater treatment is based upon various physical, chemical compounds from wastewater effluent with cleaner technology is
and biological processes. In fact, this consists of a combination of of high interest. The approach for removal of recalcitrant organic
many conventional operations such as coagulation, sedimentation, compounds is photocatalytic oxidation (Robertson, 1996). Photo-
precipitative softening, filtration, and chlorination which utilize catalytic oxidation relies upon the generation of highly reactive
potentially hazardous or polluting materials. However, some haz- short-lived hydroxyl radicals and generated holes to efficiently
ardous pollutants or chemical substances that persist in the envi- degrade or even mineralize a wide range of organics or inorganic
ronment may not be eliminated in the conventional treatment molecules (Frank and Bard, 1977). The most widely-used catalyst
process (primary, secondary and tertiary treatment) (Gaya and for water treatment applications such as pre-treated palm oil mill
Abdullah, 2008). Many of them are very slowly degradable in effluent (Cheng et al., 2016), greywater (Chong et al., 2015), removal
municipal wastewater treatment plants and they behave as of dye e.g. methylene blue (Momeni and Ghayeb, 2015) and
recalcitrant’. Bio-recalcitrant compounds, such as pharmaceuticals, Rhodamine B (Momeni and Ahadzadeh, 2016) is titanium dioxide
personal care products, surfactants, industrial chemicals, dyes, etc., (TiO2). An advancement of Ti-O nanostructures with controlled
contaminated in the environment are very dangerous even at very architectures imparting advantageous physical and chemical
properties (e.g. surface area, crystal structure, greater percentage of
active crystal facets, porosity) can broaden the applications of TiO2
on photocatalysis in higher efficiency. For example, the dominate
* Corresponding author.
E-mail address: kunlanan.kia@kmitl.ac.th (K. Kiatkittipong).
crystal facet of {010} for titanate nanoribbons calcined at 800  C can

http://dx.doi.org/10.1016/j.jclepro.2017.02.043
0959-6526/© 2017 Elsevier Ltd. All rights reserved.
906 K. Kiatkittipong, S. Assabumrungrat / Journal of Cleaner Production 148 (2017) 905e914

be beneficial for photoactivity (Kiatkittipong et al., 2011). Na2Ti6O13 2. Experimental section


nanoribbons with tunnel-like structure provided better electron/
hole separation for the water splitting reaction (Kiatkittipong et al., 2.1. Chemicals
2013). A high surface area of TiO2 nanotubes is a good nano-
structure for supporting metal to enhance the photocatalytic re- TiO2 (anatase 100%), sodium hydroxide (Ajax Finechem), hy-
action (Momeni et al., 2015). Moreover, nitrogen-doped TiO2 drochloric acid (37 vol%, (J.T.Baker)), methylene blue (MB) (Sigma-
nanotubes showed a high mobility of electron on the catalyst sur- Aldrich Chemie GmbH) and alum were used without further
face and efficient desorption of products, resulting in higher purification.
product yield rates for photocatalytic CO2 conversion (Delavari
et al., 2016). 2.2. Titanate nanotubes and titanate nanoribbons synthesis
Nanostructures such as wire, rod, tube and ribbon-shaped
structures have been synthesized by various methods such as Titanate nanostructures were synthesized using a hydrothermal
template-assisted method, electrochemical anodic oxidation procedure. Initially, 0.5 g of TiO2 and 20 mL of 10 M NaOH were
method and hydrothermal method. The hydrothermal method is stirred for 1.5 h in a plastic bottle. The slurry was transferred to a
reviewed and considered due to low cost, mass effective and high Teflon-lined autoclave, sealed and heat treated in an oven at (i)

product purity. Moreover, hydrothermal method is attractive for a 150  C to produce the nanotubes or (ii) 200 C to produce the
tight control of crystal growth and morphology (Kasuga et al., nanoribbons. The resulting suspension was washed with 0.1 M HCl
1998). In relation to photocatalysis, nanostructures with high sur- to form hydrogen titanate while sodium titanate was washed with
face area have advantages in enhancing light absorption and more water until the pH value of the wash solution reached approxi-
active site numbers. This particular aspect is important for mole- mately 6 (Kiatkittipong et al., 2011). The notation of synthesized
cules which can effectively adsorb onto the surface of the photo- photocatalyst is shown in Table 1, and the synthesis pathways used
catalyst (Hu et al., 2011). to produce photocatalysts is shown in Fig. 1.
The product obtained from alkaline hydrothermal reaction was
proposed in the form of NaxH2-xTi3O7; where 0  x  2 (Chen et al., 2.3. Characterization
2002). As alkaline hydrothermal reaction following the acid
washing was proceeded, the NaxH2-xTi3O7 was exchanged the The crystal and structural characteristics of the products were
cation with proton, forming hydrogen titanate (H2Ti3O7) (Tsai and investigated by X-ray diffractometer (XRD). The surface
Teng, 2004). As comparing the photocatalytic activity between morphology of the samples was observed by scanning electron
non-acid washed sodium titanate and acid washed titanate, the microscope (SEM) including energy dispersive X-ray spectroscopy
acid washed titanate (low Na content) was reported to promote (EDX) for measuring chemical composition, transmission electron
significantly higher activity of Brilliant-blue (KN-R) and rhodamine microscopy (TEM), and measured the specific surface area by using
6G degradation (Ribbens et al., 2008). Conversely, a higher activity the Brunauer-Emmett-Teller (BET) method. The surface bonding
for photothionine degradation was found by using NaxH2-xTi3O7. states were analyzed by X-ray photoelectron spectroscopy (XPS)
The intercalated Na atoms were suggested to trap the excited analysis.
electrons and facilitate separation of the photogenerated electrons
and holes (Pu et al., 2010). 2.4. Adsorption and photodegradation
Photocatalyst used in wastewater treatment should require post
treatment recovery, and separation of used photocatalyst from The adsorption performance and photocatalytic activity over
water before releasing is also vital. In general, colloid photo- TiO2 starting material and all synthesized samples were performed
catalysts provided a good performance in photocatalytic reaction, using aqueous MB under dark condition and UV light irradiation.
but the effectiveness of their removal from suspension is limited. 0.01 g of photocatalysts (TiO2, Na-TT, H-TT, Na-TR, H-TR) was added
TiO2 with graphite oxide has been proposed to impart an acceler- to 200 mL of MB solution. The concentrations of MB were varied in
ated sedimentation of catalyst slurry as compared to pure TiO2 a range of 10e100 of mg/L for adsorption isotherm study. The
despite the loss of catalytic activity (Szabo et al., 2013). suspension was stirred at room temperature under dark condition.
In the light of Ti-O nanostructured materials in photocatalytic The particles in sample solution were separated by micro-filter. The
applications, many studies have been developed to fabricate Ti-O MB concentrations in the sample solutions were analyzed by
nanostructured materials and to improve their properties for absorbance measurement using UVevisible spectrometer. Photo-
photocatalytic degradation. However, limited direct assessment of catalyst stability test of all synthesized samples was studied. All
Ti-O nanostructures prepared under the different hydrothermal synthesized samples were recovered from the solution by micro-
condition on adsorption and photocatalytic abilities has been filter, washed with distilled water and dried in an oven at 80  C.
attempted. Besides, the influence of morphology and phase com- The synthesized samples were then reused 3 times for MB
ponents as a photocatalyst on the separation from water before degradation.
releasing has been scarcely considered. A greater understanding of
these key aspects will be of value in developing photocatalyst and 2.5. Sedimentation
apply to industrial applications.
The key objectives deriving from this work is to develop pho- Jar test apparatus was conducted to perform the coagulation-
tocatalyst with controlled characteristics in diverse elongated
titanate nanostructures (including nanotubes and nanoribbons)
Table 1
using alkaline hydrothermal synthesis. The effects of morphologies
The notations of all synthesized photocatalyst.
and Naþ and Hþ ions intercalated in structure on adsorption and
photodegradation of methylene blue, as a model of recalcitrant Synthesized samples Notations
pollutants, are considered. Additionally, the sedimentation of tita- Sodium titanate nanotubes Na-TT
nate nanostructures as shown in a term of suspended solid in water Hydrogen titanate nanotubes H-TT
is explored. Sodium titanate nanoribbons Na-TR
Hydrogen titanate nanoribbons H-TR
K. Kiatkittipong, S. Assabumrungrat / Journal of Cleaner Production 148 (2017) 905e914 907

1. Sample Preparation 0.5 g of TiO2 + 20mL of 10 M NaOH

Nanotubes
Nanoribbons

2. Hydrothermal Treatment
Temperature 150 °C 200 °C

Time 72 h 24 h 40 h

3. Washing

0.1 M HCl Deionized water

H-TT H-TR Na-TT Na-TR

Fig. 1. Synthesis pathways used to produce sodium/hydrogen titanate nanotubes/nanoribbons.

flocculation. The 200 mL of suspension was stirred for 60 min in a where hydrogen bond array (e.g., Ti-O-H) is faster than that of the
beaker. Then, 0.5 g of alum used as a coagulant was added in the sodium bond array (e.g., Ti-O-Na) under milder conditions
suspension, and the suspension was stirred for 20 min after which (Kiatkittipong et al., 2010). However, the structures are slightly
it was allowed to rest and settle for a given period of time. After fragmented may be because Naþ is replaced by Hþ during acid
coagulation-flocculation, the particles remaining in the water were wash.
separated by filtration. The XRD patterns for all samples are shown in Fig. 3(a). 100%
anatase of TiO2 commercial used as starting material is confirmed.
3. Results and discussion With the alkaline hydrothermal reaction, the XRD pattern of Na-TT
corresponds to sodium titanate (Na2Ti3O7). The distinct peak
3.1. Structures and morphology located at 9 indicates interlayer spacing of approximately 0.8 Å.
The pattern of acid washed sample (H-TT) can be indexed to
The influences of reaction temperature and acid washing on hydrogen titanate (H2Ti3O7). The decrease in interlayer spacing to
sample morphology were investigated by SEM and corresponding 0.79 Å due to replacing Naþ ions with Hþ with a smaller ionic radius
TEM (inset) as shown in Fig. 2. Fig. 2(a) shows that morphology of compared with Naþ ions (Pu et al., 2010). The crystal structure of
starting TiO2 is non-defining morphology. After hydrothermal re- H2Ti3O7 or Na2Ti3O7 was explained to be monoclinic, where is ar-
action of starting TiO2 and NaOH at 150  C for 72 h (Na-TT), the SEM ranged in the form of TiO6 edge-sharing octahedron. Each slab is
image in Fig. 2(b) reveals that the morphology of sample is a connected at its corner to form layers that are separated by either
defined tube-like structure having an outer diameter of approxi- Naþ or Hþ ions (Kolen’ko et al., 2006). Fig. 3(a) indicates that the
mately 9 nm and length in the order of microns. Similar XRD patterns for Na-TR and H-TR are sodium titanate (Na2Ti3O7)
morphology was obtained for sample (H-TT) washed with acid after and hydrogen titanate (H2Ti3O7), respectively which are similar to
the hydrothermal condition 150  C for 72 h in Fig. 2(c). The nano- those observed for Na-TT and H-TT. The distinct peak indicating
tube diameter is similar to that observed by other (Yu et al., 2006a). interlayer spacing of nanoribbons (Na-TR) is located at 2Ө z 11. A
At higher temperature (i.e. 200  C for 40 h) (Na-TR), nanoribbon distinct difference in interlayer spacing between nanotubes and
structure having a width of approximately 10e100 nm and length nanoribbons derives from structural formation. The nanoribbon
in the order of microns is formed as shown in Fig. 2(d). At a shorter system comprises multilayered titanate sheets while bending of the
hydrothermal time of 24 h, acid washing the sample after hydro- multilayered sheets addresses the nanotube formation
thermal treatment accelerated nanoribbon formation (H-TR) in (Kleinhammes et al., 2005). Naþ and Hþ exchange process can be
Fig. 2(e). This was explained that the shorter hydrothermal time confirmed by EDX analysis as shown in Fig. 3(b) and (c). EDX
needed to produce the interlayered architecture in the acid washed analysis of the individual samples Na-TT and Na-TR showed a
908 K. Kiatkittipong, S. Assabumrungrat / Journal of Cleaner Production 148 (2017) 905e914

(a)

(b) (c)

50 nm 50 nm

(d) (e)

0.5 μm 0.5 μm

Fig. 2. SEM images of (a) TiO2 (starting material), (b) Na-TT, (c) H-TT, (d) Na-TR, (e) H-TR. Corresponding TEM (inset) showing the formation of Na-TT, H-TT, Na-TR and H-TR.

significant amount of sodium (Fig. 3(b)), while there was no sodium spectra for sodium titanate appear the binding energy of the main
ion present in both the H-TT and H-TR samples (Fig. 3(c)), indicating peaks Ti 2p, O 1s and Na 1s in Fig. 4(a). The peak of Ti 2p corre-
that sodium was successfully exchanged with Hþ. sponds to Ti 2p3/2 and Ti 2p1/2 of Ti (IV) oxidation state (Esparza
XPS spectra for the sodium titanate (Na-TT and Na-TR) and et al., 2013). The peak of O 1s attributes to Ti-O and -OH (Yu
hydrogen titanate (H-TT and H-TR) are provided in Fig. 4. The et al., 2006b). The evidence of Na 1s located at 1072 in sodium
K. Kiatkittipong, S. Assabumrungrat / Journal of Cleaner Production 148 (2017) 905e914 909

(a)

H-TR
(b)
Na-TR

H-TT

Na-TT Na-TT, Na-TR


(c)

TiO2

H-TT, H-TR

Fig. 3. (a) XRD patterns of TiO2 and all synthesized titanate, EDX analysis of (b) Na-TT, Na-TR and (c) H-TT, H-TR.

Table 2
O1S

BET surface area of TiO2 (starting material), Na-TT, H-TT, Na-


TR and H-TR.
Ti2P

Samples Surface area (m2/g)


O kill

TiO2 10
Intensity (a.u.)

C1S

Na-TT 138
H-TT 143
(b) Na-TR 32
Na1S

H-TR 35

(a) H-TR, 32e35 m2/g). When comparing the sodium titanate and
hydrogen titanate, it is apparent that the samples washed with HCl
have slightly higher surface area than the samples washed with
only water. That is, H-TT (143 m2/g) > Na-TT (138 m2/g), and H-TR
200
600

400
1,000

800

(35 m2/g) > Na-TR (32 m2/g). This may due to the minor collapse of
samples after acid washing as was illustrated in Fig. 2. Fig. 5(a)
Binding Energy (eV) shows the adsorption-desorption isotherm of Na-TT. It was
observed that the N2 isotherm of Na-TT is type IV of hysteresis loop
Fig. 4. XPS spectra of (a) Na-TT, Na-TR and (b) H-TT, H-TR. according to IUPAC classification. The hysteresis loop of Na-TT was
located at relative pressure between 0.6 and 1, where the internal
pore of tube structure was presented at P/P0 < 0.8. The hysteresis
loop at 0.8 < P/P0 < 1 corresponded to the interspace pore between
titanate was found, which agrees with the EDX analysis. The tubes. This showed good agreement with the TEM represented in
hydrogen titanate in Fig. 4(b) is also obviously observed the pres- Fig. 2. The pore distribution in Fig. 5(b) exhibits that the smaller
ence of peaks Ti 2p and O 1s. However, the peak of Na 1s in pores (pore diameter < 10 nm) corresponding to the pore inside
hydrogen titanate is not as intense as that in sodium titanate due to nanotubes, and the larger pores (pore diameter > 10 nm) attrib-
Naþ and Hþ exchange process, corresponding to EDX analysis. uting to voids occurring from overlap of nanotubes were both
Table 2 shows that the formation of nanostructures is reflected found in Na-TT. In case of nanoribbons, the adsorption-desorption
by the changes of surface area. The surface area of TiO2 (starting isotherm of Na-TR in Fig. 6(a) is a small hysteresis loop at a high
materials) was 10 m2/g. It was found that alkaline hydrothermal relative P/P0. This is due to aggregation of ribbon structure resulting
reaction can yield higher surface area products. The titanate in formation of macropores. This corresponded to the pore distri-
nanotubes (Na-TT and H-TT) have surface area of 138e143 m2/g bution in Fig. 6(b) presenting only high pore diameters, which
which is much higher than that of titanate nanoribbons (Na-TR and ascribed to the secondary pore.
910 K. Kiatkittipong, S. Assabumrungrat / Journal of Cleaner Production 148 (2017) 905e914

(a) (b)

Fig. 5. Na-TT (a) Nitrogen adsorption-desorption, (b) Pore size distribution.

(a) (b)

Fig. 6. Na-TR (a) Nitrogen adsorption-desorption, (b) Pore size distribution.

3.2. Adsorption performance

Fig. 7 shows adsorption performance of TiO2, Na-TT, H-TT, Na-TR


and H-TR in MB solution under dark condition. MB adsorption
performance appears to increase in the order: TiO2 << H-TR < Na-
TR < H-TT << Na-TT. A typical of TiO2 shows little MB adsorption,
indicating poor adsorption of MB molecules onto surface. This is
similar with the work by other researcher who reported no dark
response onto the surface of TiO2 (Cheng et al., 2014). For com-
parison, titanate nanotubes (Na-TT and H-TT) are observed to a
comparatively higher MB adsorption than titanate nanoribbons
(Na-TR and H-TR) and TiO2. These findings attributed to (1) higher
surface area providing more MB adsorbed; (2) the formation of
nanotubes having high concentration of hydroxyl groups (anionic)
forming strong electrostatic interaction between the MB (cationic)
which favor the preferential MB adsorption (Natarajan et al., 2013).
However, the difference in MB adsorption between titanate nano-
ribbons (Na-TR and H-TR) was not significant due to a small in- Fig. 7. MB adsorption under dark using of TiO2 (starting material), Na-TT, H-TT, Na-TR
crease in surface area of H-TR. The best MB adsorption was and H-TR.
provided by Na-TT, whose C/C0 could achieve 0.4 (60% adsorption)
K. Kiatkittipong, S. Assabumrungrat / Journal of Cleaner Production 148 (2017) 905e914 911

within 60 min. Na-TT adsorbs MB almost 2 times higher than H-TT Table 4
despite two samples exhibiting similar specific surface area. At this Freundlich isotherm parameters of methylene blue (MB) on TiO2 (starting material),
Na-TT, H-TT, Na-TR and H-TR.
point, it can be highlighted that Naþ ions in titanate exchanged
with a cationic dye of MB, which strongly adsorbed on the surface Photocatalyst KF nF R2
of titanate (Lee et al., 2007). This was similarly observed for titanate TiO2 starting material 2.01  102 0.52 0.985
nanoribbons, but not as much as on titanate nanotubes due to Na-TR 1.00  104 0.35 0.933
lower surface area of titanate nanoribbons. H-TR 3.04  105 0.28 0.910
Na-TT 6.97  1011 0.19 0.977
The equilibrium adsorption isotherm of MB in presence of all
H-TT 7.02  1014 0.12 0.988
samples was studied by Langmuir and Freundlich adsorption
models. The Langmuir adsorption models based on the assumption
of monolayer adsorption is expressed as
based structural materials and other classes of sorbents, the satu-
rated adsorption capacities are summarized in Table 5. The sorption
Ceq 1 Ceq performances of MB on Ti-O based material are very diverse. It
¼ þ
Qe Qm b Qm depends not only on their structures but also details of preparation
(concentration of reagent, ageing time, post treatment procedure).
where, Qe is the amount of MB adsorbed onto sample (mg/g), Ceq is Our sorption capacity in line with other researchers’ reported
MB concentration at equilibrium (mg/L), Qm is the maximum findings and the highlight is the importance of sodium incorpo-
amount of MB adsorbed per unit mass of sample (mg/g), and b is a rated in nanostructure without any further post treatment to ach-
constant related to the adsorption energy (L/mg). Table 3 shows Qm ieve high MB sorption capacity. Comparing to other classes of
and b, which were evaluated from the slope and intercept from sorbent materials, both TR and TT exhibit higher MB sorption ca-
plotting Ceq/Qe against Ceq, respectively. pacity than natural material e.g. kaolin while Na-TT is comparable
The Freundlich isotherm based on the assumption of a hetero- to bentonite and diatomaceous earth. However, the capacity is
geneous surface binding is expressed as much less than commercial activated carbon and biosorbent
(Marungrueng and Pavasant, 2007). Note that intensive review on
1 MB adsorption on natural materials, biosorbents, agricultural and
lnQe ¼ lnKF þ lnCe
nF industrial waste derived materials can be found in Rafatullah et al.
(2010).
where, KF is Freundlich constants (mg/g (L/mg)1/n) and dimen-
sionless nF suggests the favorability of adsorption. 3.3. Adsorption kinetics
Freundlich isotherm parameters of MB on different photo-
catalysts are provided in Table 4. The previous section demonstrated the highest adsorption
From Tables 3 and 4, it can be seen that the adsorption isotherms performance on Na-TT. Then, Na-TT was chosen to study adsorption
can be adequately described by Langmuir adsorption model kinetics. The adsorption kinetics of MB in presence of Na-TT was
because correlation coefficient R2 was found to be closer to unity studied by pseudo-first-order kinetic and pseudo-second-order
rather than Freundlich isotherm. This agreed with previous study kinetic. Pseudo-first-order is expressed as
(Natarajan et al., 2013). This indicates that the adsorption of MB on
TiO2 and titanate can be explained by monolayer adsorption; lnðQe  Qt Þ ¼ lnQe  k1 t
whereby a layer of MB molecules form over the surface of samples.
Pseudo second-order is expressed as
Na-TT provided the highest amounts of MB adsorption of 178.6 mg/
g with 0.39 L/mg constant related to the adsorption energy, which t 1 t
is slightly higher than the value reported by other (Xiong et al., ¼ þ
Qt k2 Qe 2 Qe
2010). The saturated adsorption capacities of MB for H-TT, Na-TR,
H-TR and TiO2 were reported as 111.1, 86.2, 80.7 and 4.7 mg/g,
where k1 is the rate constant of pseudo-first-order kinetic model
respectively. The order has been similarly observed by adsorption
(min1), k2 is rate constant of pseudo-second-order (g/mg min), Qe
performance in Fig. 7. It is worth to note that the amount of the MB
and Qt are the amounts of preferential adsorption of MB dye on Na-
adsorbed on Na-TT at equilibrium increased from 115 to 167 and
TT (mg/g) at equilibrium and at time t (min), respectively.
179 mg/g with the increase of initial MB concentration from 10 to
Table 6 showed kinetic data of the pseudo-first-order and
50 and 100 mg/L. This is due to the increase of driving force of the
pseudo-second-order kinetic models. It can be seen that the fitting
concentration gradient with increasing initial MB concentration. As
of experimental data exhibited a good compliance with pseudo-
constant sorbent loading, the removal efficiency decreased from
second-order kinetic equation, and linear plots of t/Qt versus t
59% to 17% and 9%. In order to increase removal efficiency, opti-
with correlation coefficients (R2) is closer to unity rather than
mizing amount of Na-TT loading is recommended for further
pseudo-first-order. The pseudo-second-order model can be
investigation.
explained that chemical sorption involving valency forces through
To compare the sorption performance of MB with other Ti-O
sharing or exchange of electrons between adsorbent and adsorbate
might be significant (Ho and McKay, 1999).
Table 3
Langmuir isotherm parameters of methylene blue (MB) on TiO2 (starting material), 3.4. Photocatalytic activity
Na-TT, H-TT, Na-TR and H-TR.

Photocatalyst Qm (mg/g) b (L/mg) R2 Fig. 8 presents the relative concentration of MB (C/C0) over time,
TiO2 starting material 4.7 0.06 0.999 where C and C0 are reaction concentrations at a specific time and
Na-TR 86.2 0.18 0.992 the initial concentrations of MB, respectively. Prior to photo-
H-TR 80.7 0.09 0.997 catalytic degradation, MB was adsorbed on the surface of photo-
Na-TT 178.6 0.39 0.999 catalyst for 60 min under the dark condition according to the
H-TT 111.1 0.12 0.994
aforementioned result illustrating the equilibrium adsorption
912 K. Kiatkittipong, S. Assabumrungrat / Journal of Cleaner Production 148 (2017) 905e914

Table 5
Maximum sorption capacity of MB on different Ti-O based structural materials and other class sorbents.

Sorbent Maximum sorption capacity (mg/g) References

Ti-O based materials

Sm3þ ions doped titania 2.9 Xiao et al., 2007


TiO2 nanotube 77.6 Natarajan et al., 2014
TiO2 nanotube calcined at 300  C 132 Jiang et al., 2012
Titanate/titania nanoparticles (solvothermal method) 162.2 Cheng et al., 2014
Hydrogen titanate nanowires (thin films) 80.2 Horvath et al., 2014
Hydrogen titanate nanowires (suspension) 133.9 Horvath et al., 2014
Hydrogen titanate nanoflower 147.1 Huang et al., 2012
Hydrogen titanate nanotubes 79.6 Huang et al., 2012
Hydrogen titanate nanowires 5.4 Huang et al., 2012
Sodium titanate nanotubes calcined at 400  C 133.3 Xiong et al., 2010
Sodium titanate nanotubes (Na-TT) 178.6 This study
Hydrogen titanate nanotubes (H-TT) 111.1 This study
Sodium titanate nanoribbons (Na-TR) 86.2 This study
Hydrogen titanate nanoribbons (H-TR) 80.7 This study
Other class materials
Kaolin 15.5 Ghosh and Bhattacharyya, 2002
Bentonite 150e175 Hong et al., 2009
Diatomaceous earth 198 Al-Ghouti et al., 2003
Commercial activated carbon 238, 297 Marungrueng and Pavasant (2007), Kannan and Sundaram (2001)
Macro alga Caulerpa lentillifera 417 Marungrueng and Pavasant (2007)

Table 6
Kinetic data of the pseudo-first-order and pseudo-second-order kinetic models at concentration ca. 10 mg/l.

Qe,exp (mg/g) pseudo-first-order pseudo-second-order

k1 (min1) Qe1,cal (mg/g) R2 k2 (g/(mg min)) Qe2,cal (mg/g) R2

115.0 0.079 146.9 0.938 0.075 133.3 0.973

Dark Condition UV Condition

TiO2
H-TR
Na-TR
H-TT
Na-TT

Fig. 8. Changes in the relative concentration (C/C0) of methylene blue over time by TiO2 (starting material), Na-TT, H-TT, Na-TR and H-TR.

(Fig. 6). As seen from the C/C0, a smaller degradation of MB on indicated that the photocatalytic performance relied on the corre-
typical TiO2 was obtained when compared to that of titanate sponding surface area as shown in Table 2. At this point, the pho-
nanostructures. A significant decrease in approximate 80% MB todegradation of MB on TiO2 and titanate nanostructures was found
degradation was seen for the titanate nanotubes while titanate to follow first order reaction, which was similarly observed by other
nanoribbons exhibited approximate 60% MB degradation. This (Choudhury et al., 2016). The rate constant for TiO2 was 8.56  103
K. Kiatkittipong, S. Assabumrungrat / Journal of Cleaner Production 148 (2017) 905e914 913

min1. For titanate nanotubes, the reaction rate constant (k) for Na-
TT (1.02  102 min1) was lower than that for H-TT (1.16  102
min1), despite a higher MB adsorption on Na-TT in dark condition. TiO2
The suppression of MB degradation on Na-TT may attribute that the
high amount of MB absorbed on Na-TT may block the active site for H-TR
light absorption. For nanoribbons, the reaction rates of Na-TR and
H-TR were 9.14  103 min1 and 9.32  103 min1, respectively.
Na-TR
The suppressed degradation of MB on Na-TR exhibited similar H-TT
behavior explained for titanate nanotubes, but the degradation of
MB between Na-TR and H-TR was not significantly different. This
Na-TT
behavior was also observed for oxalic degradation (Kiatkittipong
et al., 2013). This effect was not as pronounced for titanate nano-
ribbons because of their comparatively low surface area. Through
the results of this study, it was suggested that the adsorption pro-
cess and photodegradation should occur simultaneously. The
photocatalyst stability of all synthesized samples for MB degrada-
tion is illustrated in Fig. 9. The results reveal that the reaction rates
of Na-TT, H-TT, Na-TR, H-TR were not significantly different for at
least 3 cycles, indicating that all synthesized samples have photo-
catalytic stability.
Fig. 10. Particles remaining in water for suspensions over time.
3.5. Sedimentability

The sedimentability of suspension was studied by over the time immersed in HCl solution, Naþ is replaced by Hþ to form hydrogen
frame of 0e30 min. Their separability was presented in term of the titanate nanotubes (H-TT). For titanate nanoribbons, a sodium
particle remaining in water as shown in Fig. 10. Their sedimentation titanate nanoribbons (Na-TR) having surface area of 32 m2/g can be
of particles was fast, and the particles remaining in water was obtained at temperature of 200  C for 40 h. When acid washing was
below 50% during the first 5 min. It is interesting that the sedi- employed, hydrogen titanate nanoribbons (H-TR) were formed at
mentation of all synthesized titanates (particles remaining < 20%) 200  C for 24 h under hydrothermal condition. All synthesized
was more compared to TiO2 (particle remaining 30%) after 30 min. titanate nanostructures were more proficient for MB adsorption
The particles remaining in water for Na-TT and H-TT were performance as compared to starting material TiO2 due to higher
approximately 5% and 10%, respectively. Titanate nanotubes were surface areas. The largest MB adsorption capacity was seen for Na-
observed to provide a better settle down than titanate nanoribbons TT which is found as one of the highest sorption capacity among Ti-
(18% of Na-TR and 20% of H-TR remaining in water). This attributed O structural materials from the literature. That is, the adsorption of
the higher negative charge on the surface of titanate nanotubes MB dye on Na-TT was explained by Langmuir adsorption model
which is more favourable surface for a high charge density of Al3þ with a maximum amount of 178.6 mg/g. The sodium incorporation
ion to form large particles and settle as sediment. in titanate nanotubes is one of key parameters to achieve high MB
sorption capacity. The reaction of MB adsorption was pseudo-
second-order with the rate constant (k) of 0.075 g/mg min. This
4. Conclusions
was attributed to the high concentration of hydroxyl groups on the
surface, and negative charge of sodium titanate strongly adsorbed
Titanate nanostructures were synthesized by reacting anatase
cationic form of MB. Under light irradiation, the photocatalytic
TiO2 (surface area of 10 m2/g) with 10 M NaOH under hydrothermal
activity increased in the order TiO2 < titanate
condition. Sodium titanate nanotubes (Na-TT) having a surface area
nanoribbons < titanate nanotubes, which can be explained in term
of 138 m2/g were obtained at 150  C for 72 h. When Na-TT was
of their surface areas. However, the high amount of MB on sodium
titanate suppressed the reaction rate of degradation. This was
-2 -1 especially observed in the case for Na-TT due to comparatively high
X 10 (min ) surface area. That is, the reaction rate constant (k) for H-TT
(1.16  102 min1) was faster than that for Na-TT (1.02  102
min1). For sedimentability after photocatalytic reaction, the pre-
st
sent study clearly ascertained that all synthesized titanate nano-
1 cycle structures provided better separation from water with alum when
compared with TiO2. The strong interaction between titanate
nd nanotubes and the Al3þ ion leads to preferential flocculation and
k 2 cycle easy removal from the water, of which remaining Na-TT and H-TT in
water were approximately 5% and 10%, respectively.
rd
3 cycle
Acknowledgement

This work was financially supported by the Thailand Research


Fund (TRF) (Grant No. TRG5780171) and King Mongkut’s Institute of
Technology Ladkrabang. The authors thank Mr.Kasidis Naktup,
Fig. 9. Photocatalytic activity of TiO2 (starting material), Na-TT, H-TT, Na-TR and H-TR Mr.Uaychai Risri, Mr.Wongskorn Itsaraporn and Miss Araya Chue-
for MB degradation with three times of cycling uses. bupa for their assistance.
914 K. Kiatkittipong, S. Assabumrungrat / Journal of Cleaner Production 148 (2017) 905e914

References photocatalytic properties. ACS Appl. Mater. Interfaces 3, 3988e3996.


Kiatkittipong, K., Ye, C., Scott, J., Amal, R., 2010. Understanding hydrothermal tita-
nate nanoribbon formation. Cryst. Growth Des. 10, 3618e3625.
Al-Ghouti, M.A., Khraisheh, M.A.M., Allen, S.J., Ahmad, M.N., 2003. The removal of
Kleinhammes, A., Wagner, G.W., Kulkarni, H., Jia, Y., Zhang, Q., Qin, L.C., Wu, Y., 2005.
dyes from textile wastewater: a study of the physical characteristics and
Decontamination of 2-chloroethyl ethylsulfide using titanate nanoscrolls.
adsorption mechanisms of diatomaceous earth. J. Environ. Manag. 69, 229e238.
Chem. Phys. Lett. 411, 81e85.
Chen, Q., Zhou, W., Du, G.H., Peng, L.M., 2002. Trititanate nanotubes made via a
Kolen’ko, Y.V., Kovnir, K.A., Gavrilov, A.I., Garshev, A.V., Frantti, J., Lebedev, O.I.,
single alkali treatment. Adv. Mater. 14, 1208e1211.
Churagulov, B.R., Van Tendeloo, G., Yoshimura, M., 2006. Hydrothermal syn-
Cheng, C.K., Deraman, M.R., Ng, K.H., Khan, M.R., 2016. Preparation of titania doped
thesis and characterization of nanorods of various titanates and titanium di-
argentum photocatalyst and its photoactivity towards palm oil mill effluent
oxide. J. Phys. Chem. B 110, 4030e4038.
degradation. J. Clean. Prod. 112, 1128e1135.
Lee, C.K., Liu, S.S., Juang, L.C., Wang, C.C., Lyu, M.D., Hung, S.H., 2007. Application of
Cheng, Y.H., Gong, D., Tang, Y., Ho, J.W.C., Tay, Y.Y., Lau, W.S., Wijaya, O., Lim, J.,
titanate nanotubes for dyes adsorptive removal from aqueous solution.
Chen, Z., 2014. One-pot solvothermal synthesis of dual-phase titanate/titania
J. Hazard. Mater. 148, 756e760.
Nanoparticles and their adsorption and photocatalytic performances. J. Solid
Marungrueng, K., Pavasant, P., 2007. High performance biosorbent (Caulerpa len-
State Chem. 214, 67e73.
tillifera) for basic dye removal. Bioresour. Technol. 98, 1567e1572.
Chong, M.N., Cho, Y.J., Poh, P.E., Jin, B., 2015. Evaluation of Titanium dioxide pho-
Momeni, M.M., Ghayeb, Y., 2015. Synthesis and characterization of iron-doped
tocatalytic technology for the treatment of reactive Black 5 dye in synthetic and
titania nanohoneycomb and nanoporous semiconductors by electrochemical
real greywater effluents. J. Clean. Prod. 89, 196e202.
anodizing method as good visible light active photocatalysts. J. Mater. Sci.
Choudhury, B., Bayan, S., Choudhury, A., Chakraborty, P., 2016. Narrowing of band
Mater. Electron. 26, 5509e5517.
gap and effective charge carrier separation in oxygen deficient TiO2 nanotubes
Momeni, M.M., Ghayeb, Y., Ghonchegi, Z., 2015. Fabrication and characterization of
with improved visible light photocatalytic activity. J. Colloid Interface Sci. 465,
copper doped TiO2 nanotube arrays by in situ electrochemical method as effi-
1e10.
cient visible-light photocatalyst. Ceram. Int. 41, 8735e8741.
Delavari, S., Amin, N.A.S., Ghaedi, M., 2016. Photocatalytic conversion and kinetic
Momeni, M.M., Ahadzadeh, I., 2016. Copper photodeposition on titania nanotube
study of CO2 and CH4 over nitrogen-doped titania nanotube arrays. J. Clean.
arrays and study of their optical and photocatalytic properties. Mater. Res.
Prod. 111, 143e154.

ndez, T., Borges, M.E., Alvarez-Galv n, M.C., Ruiz-Morales, J.C., Innov. 20, 44e50.
Esparza, P., Herna a
Natarajan, T.S., Natarajan, K., Bajaj, H.C., Tayade, R.J., 2013. Enhanced photocatalytic
Fierro, J.L.G., 2013. TiO2 modifications by hydrothermal treatment and doping to
activity of bismuth-doped TiO2 nanotubes under direct sunlight irradiation for
improve its photocatalytic behaviour under visible light. Catal. Today 210,
degradation of Rhodamine B dye. J. Nanoparticle Res. 15, 1669e1687.
135e141.
Natarajan, T.S., Bajaj, H.C., Tayade, R.J., 2014. Preferential adsorption behavior of
Frank, S.N., Bard, A.J., 1977. Heterogeneous photocatalytic oxidation of cyanide ion
methylene blue dye onto surface hydroxyl group enriched TiO2 nanotube and
in aqueous solutions at TiO2 powder. J. Am. Chem. Soc. 99, 303e304.
its photocatalytic regeneration. J. Colloid Interface Sci. 433, 104e114.
Gaya, U.I., Abdullah, A.H., 2008. Heterogeneous photocatalytic degradation of
Pu, Y.C., Chen, Y.C., Hsu, Y.J., 2010. Au-decorated NaxH2-xTi3O7 nanobelts exhibiting
organic contaminants over titanium dioxide: a review of fundamentals, prog-
remarkable photocatalytic properties under visible-light illumination. Appl.
ress and problems. J. Photochem. Photobiol. C Photochem. Rev. 9, 1e12.
Catal. B Environ. 97, 389e397.
Ghosh, D., Bhattacharyya, K.G., 2002. Adsorption of methylene blue on kaolinite.
Ribbens, S., Meynen, V., Tendeloo, G.V., Ke, X., Mertens, M., Maes, B.U.W., Cool, P.,
Appl. Clay Sci. 20, 295e300.
Vansant, E.F., 2008. Development of photocatalytic efficient Ti-based nanotubes
Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes. Pro-
and nanoribbons by conventional and microwave assisted synthesis strategies.
cess Biochem. 34, 451e465.
Microporous Mesoporous Mater. 114, 401e409.
Hong, S., Wen, C., He, J., Gan, F., Ho, Y.S., 2009. Adsorption thermodynamics of
Robertson, P.K., 1996. Semiconductor photocatalysis: an environmentally acceptable
methylene blue onto bentonite. J. Hazard. Mater. 167, 630e633.
alternative production technique and effluent treatment process. J. Clean. Prod.
Horv gyi, I., Forro
ath, E., Szila , L., Magrez, A., 2014. Probing titanate nanowire surface
4, 203e212.
acidity through methylene blue adsorption in colloidal suspension and on thin
Rafatullah, M., Sulaiman, O., Hashim, R., Ahmad, A., 2010. Adsorption of methylene
films. J. Colloid Interface Sci. 416, 190e197.
blue on low-cost adsorbents: a review. J. Hazard. Mater. 177, 70e80.
Hu, K., Xiao, X., Cao, X., Hao, R., Zuo, X., Zhang, X., Nan, J., 2011. Adsorptive sepa-
Szabo, T., Veres, T., Cho, E., Khim, J., Varga, N., Dekany, I., 2013. Photocatalyst sep-
ration and photocatalytic degradation of methylene blue dye on titanate
aration from aqueous dispersion using graphene oxide/TiO2 nanocomposites.
nanotube powders prepared by hydrothermal process using metal Ti particles
Colloids Surfaces A Physicochem. Eng. Asp. 433, 230e239.
as a precursor. J. Hazard. Mater. 192, 514e520.
Tsai, C.C., Teng, H., 2004. Regulation of the physical characteristics of titania
Huang, J., Cao, Y., Liu, Z., Deng, Z., Wang, W., 2012. Application of titanate nano-
nanotube aggregates synthesized from hydrothermal treatment. Chem. Mater.
flowers for dye removal: a comparative study with titanate nanotubes and
16, 4352e4358.
nanowires. Chem. Eng. J. 191, 38e44.
Xiao, Q., Si, Z.-C., Zhang, J., Xiao, C., Yu, Z., Qiu, G.-Z., 2007. Effects of samarium
Jiang, F., Zheng, S., An, L., Chen, H., 2012. Effect of calcination temperature on the
dopant on photocatalytic activity of TiO2 nanocrystallite for methylene blue
adsorption and photocatalytic activity of hydrothermally synthesized TiO2
degradation. J. Mater. Sci. 42, 9194e9199.
nanotubes. Appl. Surf. Sci. 258, 7188e7194.
Xiong, L., Yang, Y., Mai, J., Sun, W., Zhang, C., Wei, D., Chen, Q., Ni, J., 2010. Adsorption
Kannan, N., Sundaram, M.M., 2001. Kinetics and mechanism of removal of methy-
behavior of methylene blue onto titanate nanotubes. Chem. Eng. J. 156,
lene blue by adsorption on various carbons-a comparative study. Dyes Pigments
313e320.
51, 25e40.
Yu, J., Yu, H., Cheng, B., Trapalis, C., 2006a. Effects of calcination temperature on the
Kasuga, T., Hiramatsu, M., Hoson, A., Sekino, T., Niihara, K., 1998. Formation of ti-
microstructures and photocatalytic activity of titanate nanotubes. J. Mol. Catal.
tanium oxide nanotube. Langmuir 14, 3160e3163.
A Chem. 249, 135e142.
Kiatkittipong, K., Iwase, A., Scott, J., Amal, R., 2013. Photocatalysis of heat treated
Yu, J., Yu, H., Cheng, B., Zhou, M., Zhao, X., 2006b. Enhanced photocatalytic activity
sodium- and hydrogen-titanate nanoribbons for water splitting, H2/O2 gener-
of TiO2 powder (P25) by hydrothermal treatment. J. Mol. Catal. A Chem. 253,
ation and oxalic acid oxidation. Chem. Eng. Sci. 93, 341e349.
112e118.
Kiatkittipong, K., Scott, J., Amal, R., 2011. Hydrothermally synthesized titanate
nanostructures: impact of heat treatment on particle characteristics and

You might also like