You are on page 1of 8

G Model

JIEC-5562; No. of Pages 8

Journal of Industrial and Engineering Chemistry xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


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

Synergetic photocatalytic-activity enhancement of lanthanum doped


TiO2 on halloysite nanocomposites for degradation of organic dye
Jewon Leea , Sicheon Seongb , Soyeong Jinb,c , Youngdo Jeongc,d,* , Jaegeun Nohb,e,**
a
Department of Convergence of Nanoscience, Hanyang University, 222 Wangsimni-ro, Seongdong-gu, Seoul 04763, Republic of Korea
b
Department of Chemistry, Hanyang University, 222 Wangsimni-ro, Seongdong-gu, Seoul 04763, Republic of Korea
c
Center for Biomaterials, Biomedical Research Institute, Korea Institute of Science and Technology (KIST), Seoul 02792, Republic of Korea
d
Department of HY-KIST Bio-convergence, Hanyang University, 222 Wangsimni-ro, Seongdong-gu, Seoul 04763, Republic of Korea
e
Institute of Nano Science and Technology, Hanyang University, 222 Wangsimni-ro, Seongdong-gu, Seoul 04763, Republic of Korea

A R T I C L E I N F O A B S T R A C T

Article history: A reusable and efficient photo-catalyst based on halloysites (HNTs) and La3+ ion-doped TiO2
Received 18 March 2021 nanoparticles was prepared using a sol-gel process and a sonochemical method for decomposing
Received in revised form 17 May 2021 organic dyes. As a support, the HNTs increased the adsorption efficiency of substrates, improving the
Accepted 20 May 2021
photocatalytic activity of TiO2 nanoparticles by 3 times. The La3+ ion doping of TiO2 decreased the
Available online xxx
bandgap from 3.25 eV to 3.01 eV, expanding the spectral response to the visible region and enhancing the
photocatalytic activity under the sunlight by 1.5 times. Through the integration of the halloysite support
Keywords:
and the La3+ ion doping of TiO2, the synergistic enhancement of photocatalytic activity was observed and
Halloysite
Nanotubes
systematically investigated using XRD, Raman spectroscopy, FT-IR, FE-SEM, EDX-TEM, and zeta potential.
TiO2 nanoparticle In this research, we used Rhodamine B molecules as a model pollutant. During the decomposing process,
Nanocomposite the HNTs enhanced the local dye concentration nearby the photo-catalyst through the electrostatic
Photodegradation interaction, thereby boosting the photocatalytic activity by 3 times. Additionally, the catalyst can be
Re-usability separated by simple centrifugation and reused at least five recycles without any activity decrease. The
halloysite support, the La3+ ion doping of TiO2 and their simple synthesis approach could provide a
practical tool for environmental purification in wastewater using sunlight.
© 2021 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

Introduction also quantum confinement and high surface to volume ratio from
their nano-size [13,14]. However, the feature that TiO2 nano-
Textile industries produce large amounts of wastewater, of particles tend to aggregates in the solution can cancel the benefit
which organic dyes are toxic molecules in the aquatic environment from the size, and their large energy bandgap of 3.2 eV limits the
[1–3]. Although many strategies such as filtration and adsorption catalysts to possess the high activity only in UV light, not the
have been suggested to remove pollutants from wastewater, a cost- sunlight [15].
effective and eco-friendly method is still required to purify Immobilization of TiO2 at supporting materials can prevent
enormous amounts of wastewater [4–10]. The use of photo- aggregation, sustaining their catalytic activity [16]. As a support
catalysts that can produce reactive oxygen radicals can be a candidate, halloysite nanotubes (HNTs) have advantages such as
practical candidate since the sunlight can provide the energy eco-friendly properties and biocompatibility with low cytotoxicity
needed for the chemical reaction without any cost [11,12]. Among [17–24]. The HNTs, natural porous materials, are two-layered
many photocatalysts, anatase TiO2 nanoparticles have been aluminosilicate clay and have exhibited adsorption ability for
focused due to not only their high activity from TiO2 itself but organic pollutants due to their inherent hollow nanotube structure
and different outside and inside surficial properties [25–28]. Since
the inside surface of the HNTs consists of an AlOH structure with
* Corresponding author at: Center for Biomaterials, Biomedical Research positive charges and the outside surface possess a Si–O–Si
Institute, Korea Institute of Science and Technology (KIST), Seoul, 02792, Republic structure with negative charges, both charged pollutants can
of Korea. adsorb on their surface [29–32]. Lazzara et al. synthesized the TiO2
** Corresponding author at: Department of Chemistry, Hanyang University, 222
deposited on HNT, showing their stability in water for a month and
Wangsimni-ro, Seongdong-gu, Seoul 04763, Republic of Korea.
E-mail addresses: zerodegree@kist.re.kr (Y. Jeong), jgnoh@hanyang.ac.kr (J. Noh). the enhanced activity by 1.75 times [33]. However, the effect of

https://doi.org/10.1016/j.jiec.2021.05.029
1226-086X/© 2021 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.

Please cite this article as: J. Lee, S. Seong, S. Jin et al., Synergetic photocatalytic-activity enhancement of lanthanum doped TiO2 on halloysite
nanocomposites for degradation of organic dye, J. Ind. Eng. Chem., https://doi.org/10.1016/j.jiec.2021.05.029
G Model
JIEC-5562; No. of Pages 8

J. Lee, S. Seong, S. Jin et al. Journal of Industrial and Engineering Chemistry xxx (xxxx) xxx–xxx

the increased adsorption efficiency caused by the HNTs on the Materials and methods
photocatalytic activity was not systematically studied, the TiO2
used in their study still has a large bandgap, and the catalytic Preparation of TiO2 @HNTs composite
activity is still needed to improve more [34].
On the other hand, to reduce the bandgap energy of TiO2, TiO2 nanoparticle on the HNTs surface was synthesized by the
doping TiO2 with metal /nonmetal atoms has received huge sol–gel method [46]. To 24 mL of isopropanol, 5.04 mL of TTIP was
attention [35–43]. Asahi et al. and Khan et al. reported that doping added in a 150 mL round-bottom flask. A mixture of 16 mL of
TiO2 with nitrogen or carbon could lower its bandgap and shift its isopropyl alcohol, 16 mL of water, and 0.2 mL of nitric acid was
optical response to the visible region [5,44]. These doped TiO2 added dropwise to the reaction mixture with vigorous stirring. The
nanoparticles showed the enhanced catalytic activity and lowered resultant reaction mixture was continuously stirred for 2 h, 1.3 g of
bandgap, providing catalytic activity under visible irradiation [45]. HNT was added, and stirred for 2 h, then the mixture was aged at
However, to date, the systematic study for the synergistic effect of room temperature for 24 h. The reaction mixture was centrifuged
the deposition of doping TiO2 on HNTs on photocatalytic activity is at 8000 rpm for 10 min. The separated catalyst composite was
lacking. Additionally, to practically use the metal/nonmetal doped washed three times with a mixture of ethanol and water and dried
TiO2@HNTs catalysts, a simple and effective synthesis protocol is in a vacuum oven at 85  C for 12 h. The dried solid catalyst was
required. ground to obtain a fine powder and calcined at 350  C for 2 h under
Here, we report a simple approach to synthesize La3+ doped the ambient condition at a heating rate of 10  C/min. (Fig. 1)
TiO2 on HNTs (La/TiO2@HNTs) using a sol-gel process and a Comparative catalyst TiO2 nanoparticles were synthesized in the
sonochemical method, and also, the synergistic enhancement of same process except for HNTs. The formation process of
photocatalytic activity. First, TiO2 was synthesized by a low- immobilization TiO2 on HNTs caused by the involved two
temperature sol–gel method using titanium(IV) isopropoxide reactions: (1) hydrolysis of the precursor in the acidic or basic
precursor on the HNT surface. Then, La3+ ions were doped with mediums and (2) polycondensation of the hydrolyzed products.
lanthanum nitrate hexahydrate precursor by a sonochemical The titanium precursors were hydrolyzed and the formed the
method (Fig. 1). We analyzed the catalysts' structure using XRD, nanoparticle as a sol. Through the gelation of the sol, the
Raman spectroscopy, FT-IR, FE-SEM, EDX-TEM, and zeta potential. nanoparticles attached on the surface of HNTs were growing.
The fabricated catalyst could decompose Rhodamine B molecules
as a model pollutant. During the process, the HNTs increased the Preparation of lanthanum doped TiO2@HNTs (La/TiO2@HNTs)
local concentration of the dye nearby TiO2 through the electro- composite
static interaction, thereby improving the photocatalytic activity of
TiO2 nanoparticles by 3 times, showing the synergistic enhance- The La-doped TiO2@HNTs composite was prepared by an
ment of photocatalytic activity. Analysis of the Brunauer-Emmett- ultrasound-assisted wet impregnation method [47]. The sono-
Teller (BET) surface area and Barrett-Joyner-Halenda (BJH) pore chemical synthesis method is based on the acoustic cavitation
size distribution demonstrated that the porous structure of HNTs phenomenon. As the bubbles grow and collapse by ultrasonic
could induce adsorption between the photocatalyst and the dye. waves, the local micro-environment with high temperature and
Through the kinetic study, we could demonstrate that the high pressure is formed instantaneously, resulting in uniform
adsorption process was a pseudo-second-order and the decompo- nanoparticles and bonding between the dopant and TiO2 [48]. For
sition process was a pseudo-first-order. The La3+ ion doping of TiO2 the synthesis, 0.1590 g of La(NO3)36H2O (98%, Junsei) and 1.0 g of
decreased the bandgap from 3.25 eV to 3.01 eV, expanding the TiO2@HNTs were dispersed in 100 mL of distilled water and stirred
spectral response to the visible region and enhancing the for 30 min. The amount of the La ion for the photo-catalysts
photocatalytic activity under the sunlight by 1.5 times. Addition- synthesis was optimized using photo-catalytic activity of Rhoda-
ally, using simple centrifugation, we separated the catalysts from mine B (Fig. S1). Then, the solution was ultrasonically processed for
the reaction mixture and reused them. In five-times recycles, we 2 h (40 kHz, Branson 5510). The reaction mixture was centrifuged
could not observe any catalytic activity decrease (Scheme 1). at 8000 rpm for 10 min and the obtained product dried in a vacuum
oven at 85  C for 12 h. (Fig. 1) Comparative catalysts, La/TiO2 and

Scheme 1. Design of La/TiO2@HNTs and their synergistic activity enhancement. The


HNTs can improve the absorption efficiency between the organic pollutant and the
Fig. 1. Synthesis processes of La/TiO2@HNTs nanocomposites: TiO2 nanoparticle on photocatalysts. The lanthanum doping reduces the bandgap of TiO2, allowing the
the HNTs surface was synthesized using the sol–gel method, and the La-doped photolysis of organic dyes under sunlight. The combination of the HNTs support and
TiO2@HNTs composite was prepared by an ultrasound-assisted wet impregnation the lanthanum-doping synergistically enhanced the photocatalytic activity of the
method. catalysts.

2
G Model
JIEC-5562; No. of Pages 8

J. Lee, S. Seong, S. Jin et al. Journal of Industrial and Engineering Chemistry xxx (xxxx) xxx–xxx

La@HNTs were synthesized using TiO2 nanoparticles and HNTs, to the dye solutions at concentrations of 2.5, 5.0, and 10 mg/L,
respectively, in the same process. respectively. During magnetic stirring at 800 rpm for 1 h, the
absorbance at 554 nm of the rhodamine B solution was obtained by
Structural analysis of La/TiO2@HNTs UV–vis analysis according to the interval. The removal efficiency,
the amount of dye adsorbed at time qt, at equilibrium qe, and the
The crystal structure was investigated with high-resolution pseudo-second-order rate constants (k2) were calculated from the
X-ray diffraction (XRD) Bruker, D-8 Discover equipment (40 kV, equation below [49].
40 mA, Cu K-alpha (l = 1.5406 Å), scanning from 5 to 75 ). To obtain
C0  Ce
FT-IR spectrum and Raman spectrum for the samples, FTLA 2000 Removal efficiency ð%Þ  ¼   100 ð1Þ
C0
(BK Instruments Inc.) and Renishaw inVia FSM (laser wavelength =
532 nm) were used, respectively. For SEM and TEM images, we
used S-4800 (Hitachi) and JEM 2100F (JEOL). The BET surface area
ðC0  Ce ÞV
was measured using advanced free space measurement, Belsorp- qe  ðmg=gÞ  ¼   ð2Þ
m
mini II, MicrotracBEL Corp. The zeta potential of the sample was
measured using a Zetasizer Nano ZS (Malvern Instruments Ltd.).
For the UV–vis measurement, Evolution 60S (Thermo Fisher ðC0  Ct ÞV
Scientific) was used. qt  ðmg=gÞ  ¼   ð3Þ
m

Measurement of rhodamine B adsorption


t 1 1
Adsorption experiment according to the initial concentration of ¼ þ t ð4Þ
qt k2 qe 2 qe
the dye was evaluated by adding 0.1 g of La/TiO2@HNTs composite

Fig. 2. (a) XRD patterns, (b) Raman spectra, and (c) FT-IR spectra of TiO2, La/TiO2, HNTs, La@HNTs, TiO2@HNTs, and La/TiO2@HNTs. FE-SEM images of (d) TiO2, (e) La/TiO2, f)
HNTs (g) La@HNTs, (h) TiO2@HNTs, and (i) La/TiO2@HNTs. (j) Transmission electron microscope (TEM) image of La/TiO2@HNTs. (k) Energy dispersive spectroscopy (EDS)
mapping for Si, Al, O, and Ti elements. (l) EDS mapping spectra.

3
G Model
JIEC-5562; No. of Pages 8

J. Lee, S. Seong, S. Jin et al. Journal of Industrial and Engineering Chemistry xxx (xxxx) xxx–xxx

C0 and Ct are concentration at initial time and time t, formation of nanocomposites, we also performed the Raman
respectively. m is the weight of catalyst (unit: g), and V is the analysis (Fig. 2b). The TiO2 anatase peaks appeared in the
volume of dye solution (unit: L). qe and qt are the rhodamine B spectrum, with a strong peak at 145 cm1 and low peaks at 399,
adsorption capacity (unit: mg/g) and time (unit: min) at 516, and 640 cm1 [54]. FT-IR was also measured by making KBR
equilibrium, respectively. k2 is the rate constant of the pseudo- pellets with catalyst powder. (Fig. 2c) Various peaks were assigned
second-order (unit: g/mg min). For the rhodamine B adsorption in the plot (Fig. 2c) (The broad peak at 516 cm1: TiO bending
experiment the catalyst (0.02, 0.05, 0.10, 0.15, and 0.20 g) was [55], the double peaks at 3695 and 3625 cm1: AlOH stretching
added to 20 mL dye solutions with 5.0 mg/L concentration. The the inner surface of HNTs, the 1034 cm1: Si O bonds, the
change of adsorption depending on pH was tracked (pH 2.0, 4.0, 910 cm1: Al OH, the 690 cm1: low stretching band of Si O, and
6.0, 8.0, and 10) by adding 0.1 g of catalyst to 20 mL of the the 540 cm1: the vibration of Al O Si. [56]). To check the size
rhodamine B solution (5 mg/L concentration). and shape of the nanocomposite, we obtained the FE-SEM images
of each sample. The image of TiO2 (Figs. 2d, S2a) and La/TiO2
Photodegradation of rhodamine B using the nanocomposites (Figs. 2e, S2b) showed the aggregated TiO2 particles with 15 nm-
size. In the HNTs (Figs. 2f, S1c) and the La@HNTs (Figs. 2g, S3a), the
For the photocatalytic dye degradation of rhodamine B smooth surfaces of the HNTs and their long and pipe-type structure
according to the initial dye concentration, 0.02 g of La/TiO2@HNTs were observed, revealing the La3+ ion were unable to adsorb the
was added to each solution (20 mL, concentrations: 2.5, 5, and HNTs’ surface. Through the SEM image of TiO2@HNTs (Figs. 2h,
10 mg/L), and then stirred in a dark place for 30 min for sufficient S3b) and La/TiO2@HNTs (Figs. 2i, S3c), we could confirm the TiO2
adsorption. After adsorption proceeds, The UV lamp (Ushio-SP9, nanoparticles were attached to HNTs. In the La/TiO2@HNTs images,
250 W, 365 nm) was irradiated while stirring continuously, and the the bright spot of TiO2 nanoparticle was observed, presumably
absorbance of the dye decomposed over time was measured using indicating the lanthanum doped area caused by the high-electron
a UV–vis spectrometer. The dye degradation efficiency was density of the Lanthanum. The TEM image of the La/ TiO2@HNTs
calculated using the equation below and the rate constant using sample also showed the TiO2 nanoparticles on the surface of HNTs
the pseudo-first-order kinetic model below [50]. with a 15 nm lumen (Figs. 2j, S4–S6). Through the TEM-EDS
mapping, we could check the existence of Si, Al, O, and Ti atoms on
C0  C
Degradation efficiency ð%Þ  ¼     100 ð5Þ the surface of HNTs (Figs. 2k, S7). The EDS map spectrum (Figs. 2l,
C0
S8) showed the Al, Si, O, Ti, and La atoms. Since the EDS mapping
can detect the lanthanum atoms but cannot quantify them, we
performed the ICP-MS to find the composition of titanium and
C0
ln ¼ k1 t ð6Þ lanthanum on the La/TiO2@HNTs. The catalysts consisted of 58% of
C
TiO2 nanoparticles and 42% of HNTs. The lanthanum doping was
C0 and C are initial and decomposed concentrations of 0.6% based on the TiO2 (Fig. S8, Table S1).
rhodamine B (unit: mg/L), respectively. k1 is the rate constant
of the pseudo-first-order (unit: min1). The kinetic study in terms Porous properties and interfacial charge of the La/TiO2@HNTs
of the amount of the catalyst was performed under UV conditions composite
by adding 0.01, 0.02, and 0.04 g of the catalyst to a 20 mL solution of
rhodamine B (concentration: 5 mg/L). Rhodamine B degradation The surface area, pores, and surface charge of the nano-
depending on pH was evaluated by individually adding 0.02 g of composite take a critical role in regulating the accessibility of
catalyst to the dye solution at pH 2.0, 6.0, and 10. For various substrates, dictating the catalytic activity. To characterize the
catalyst comparison experiments, TiO2, La/TiO2, La@HNTs, nanocomposites' porous structure, we measured the nitrogen
TiO2@HNT, and La/TiO2@HNTs composites were added to 20 mL adsorption and desorption isotherms and pore size distributions of
of rhodamine B solution (concentration: 5 mg/L), and stirred in the the photocatalysts. In the BET plot, like the HNTs showed the
dark for 30 min. After continuous stirring, irradiation using a UV mesoporous characteristics, the composite also displayed similar
lamp, recoding the absorbance change of the dye decomposing patterns. Through the BJH porosity measurements, we could attain
over time, the removal efficiency, and decomposition rate the composites' pore size and pore volume (Table 1). From the BET
constants were calculated. Dye decomposition experiments in plot, we could interpret that the adsorption and desorption
the sunlight were evaluated using direct sunlight from 2:00 pm to isotherm of the HNTs and TiO2 is a type IV with H3 wedged hole
5:30 pm. (September in South Korea). [57] and a type IV with an H2 bottleneck hole, respectively [58].
From the adsorption and desorption isotherm of TiO2@HNTs
Results and discussion composite, we discovered that the pore volume and the surface
area increased. The La/TiO2@HNTs composites exhibited the
Surface morphology and chemical composition analysis largest BET surface area of 82.844 m2/g. The BJH pore radius of
the composite was decreased from 6.95 nm for HNTs to 3.53 nm,
To investigate the microstructure of TiO2 and the HNTs, we indicating the TiO2 nanoparticles on the surface HNTs block the
scanned the sample powders calcined at 350  C using XRD (Fig. 2a).
In the spectrum of the La/TiO2@HNTs samples, seven peaks were Table 1
observed. Among them, the five peaks (25.3 , 37.8 , 48.00 , 54.5 , BET surface area and BJH porosity measurements.
and 62.6 indexed TiO2-anatase phase (101), (004), (200), (211),
Sample BET surface BJH pore BJH pore
and (204), respectively) were well-matched with the spectrum of area [m2/g]a volume [cm3/g]a radius [nm]a
the TiO2 sample, indicating the TiO2 nanoparticles on the
TiO2 67.354 0.1069 2.71
La/TiO2@HNTs had the anatase phase. The other two peaks La/TiO2 76.927 0.1324 2.71
(11.8 and 19.9 ) could be assigned to (001), (100) plane of the HNTs 40.590 0.1998 6.95
HNTs. As compared to standard diffraction data for the TiO2 La@HNTs 39.537 0.2871 6.95
anatase phase in the previous report, the peak positions and TiO2@HNTs 78.404 0.2160 3.53
La/TiO2@HNTs 82.844 0.1990 3.53
relative intensities of the diffraction lines were well-paired
[51–53]. To underpin the structure of TiO2@HNTs and the a
Calculated form the BET plot.

4
G Model
JIEC-5562; No. of Pages 8

J. Lee, S. Seong, S. Jin et al. Journal of Industrial and Engineering Chemistry xxx (xxxx) xxx–xxx

Fig. 4. (a) UV absorption spectra of TiO2 and La/TiO2. (b) Calculated Tauc plot for
band gap energy of TiO2 and La/TiO2.

Fig. 3. (a) N2 adsorption–desorption isotherms. (b) BJH pore size distributions. (c)
Zeta potential depending on pH change/(d) Zeta potential value at pH 6.0.

pore. The interfacial charge of the composites was measured by


using the zeta potential. The zeta potential value of HNTs and TiO2
were –34.8 mV, 2.7 mV at pH 6.0, respectively. The zeta potential
of TiO2@HNT appeared in the form of a combination of HNTs and
TiO2, and the lanthanum doping rendered for the zeta potential
value of the catalyst surface to move in a positive direction,
showing the less negative charge (Fig. 3).

Bandgap decrease of the La ion doping on TiO2 nanoparticles

To measure the bandgap of TiO2 and La/TiO2, as shown in


Fig. 5. (a) Effect of initial concentration for rhodamine B adsorption onto La/
Fig. 4a–b, we obtained the UV–vis absorption spectrum and TiO2@HNTs. (b) Pseudo-second-order kinetics for rhodamine B adsorption. (c) Effect
transferred them to the Tauc-plot TiO2 and the La/TiO2 (Fig. 4). The of adsorbent dose. (d) Effects of initial solution pH on adsorption.
Tauc-plot was created by (Ahy)2 versus hy (A: absorbance, h:
Plank’s constant, and y: frequency of light) [59]. We inferred that adsorption efficiency in terms of pH change. The adsorption
the La ion doping decreases the bandgap of TiO2, broadening the removal efficiency increased up to 82.23% at pH 2.0 since the
active wavelength range. Compared to pure TiO2, the absorption of zwitterionic form of rhodamine B changes to the cationic form in
lanthanum-doped TiO2 extends into the visible region. In the Tauc- acidic conditions and is more adsorbed on the negatively charged
plot (Fig.4b), the bandgap of TiO2 and La/TiO2 showed 3.25 and catalyst surface (Fig. 5d).
3.01 eV, respectively. Thus, the extended absorption range
indicates that the La ion-doped photocatalysts can be activated Photocatalytic degradation of rhodamine B using the La/TiO2@HNTs
by visible light as well as UV light. After the immobilization of the
TiO2 and La/TiO2 on HNTs, we also measured and calculated the Encouraged by the high adsorption efficiency of the La/
bandgap change (Fig. S9) In the Tauc-plot (Fig. S9b), the bandgap of TiO2@HNTs, we next monitored the photolysis of rhodamine B
TiO2@HNT and La/TiO2@HNT were 3.18 and 3.01 eV, respectively. under UV conditions depending on the initial dye concentration,
We assumed that the bandgap difference between before and after catalyst dosing, and pH condition. In the condition of the low dye
immobilization caused by the absorption property of HNT in the concentration (2.5 mg/L), 95.34% of dyes were decomposed in
TiO2 absorption range. 10 min (Figs. 6a, b), and the degradation efficiency increased as the
dosing of the catalyst increased (Figs. 6c, d). The pH change
Adsorption of rhodamine B using the La/TiO2@HNTs affected the adsorption of dyes and the degradation efficiency, only
under less than pH 6 (Figs. 6e, f). Despite only the 15% adsorption
Under the dark condition for preventing the photodegrade of difference between the pH 6 and the pH 2 conditions (Fig. 5d), the
dyes, we evaluated the dye (rhodamine B) adsorption of the rate constant increases by more than 50% (Fig. 6f), indicating the
La/TiO2@HNTs. As shown in Fig. 5a, over 90% of dyes could adsorb synergistic activity enhancement caused by the adsorption
on the catalysts in 10 min regardless of the dye concentration. As efficiency increase.
the initial dye concentration increased, the amount of adsorbed Next, we monitored the photodegradation of rhodamine B
dyes increased (Fig. 5b and Table S2), and the adsorption curve under UV using the catalysts and their precursors (Fig. 7a, c). The
matched the pseudo-second-order kinetic model [60]. The rhodamine B was transparently decomposed when the
adsorption efficiency of dye increases as the catalyst dosing La/TiO2@HNTs were added under UV in 20 min, showing the
increases (Fig. 5c), implying the nanocomposites' adsorption 99.36% decomposition yield. As shown in Fig. 7e and f,
ability to the targeted dyes [61,62]. Additionally, we tracked the the La/TiO2@HNTs displayed the pseudo-first-order kinetics for

5
G Model
JIEC-5562; No. of Pages 8

J. Lee, S. Seong, S. Jin et al. Journal of Industrial and Engineering Chemistry xxx (xxxx) xxx–xxx

Fig. 6. Using La/TiO2@HNTs, rhodamine B photodegradation (a) and its kinetics (b) depending on initial dye concentration. The photodegradation (c) and its kinetics (d)
depending on the catalysts dosing. (e) The photodegradation (e) and its kinetics (f) depending on the solution pH.

the rhodamine B degradation (Table S3). Although there is no recycles, demonstrating reusability and its simple separation
activity difference between La/TiO2 and TiO2 under UV light method (Figs. 8a, S10).
because UV light can provide enough photon energy for electrons
to excite the bandgap, however, the La/TiO2@HNTs showed a much Active species trapping and mechanism study
higher rate constant than TiO2/HNTs (Fig. 7e). These results
indicate that lanthanum doping can reduce the bandgap of TiO2 Tracking of the active species in the photocatalytic process was
and improve the reaction rate through the combination of the carried out through trapping experiments. Benzoquinone (BQ),
HNTs support, implying the synergistic activity enhancement. We triethanolamine (TEOA), and isopropyl alcohol (IPA) scavenge
 
also monitored the photodegradation under sunlight to demon- O2–, h+ (hole), and OH, respectively [64,65]. Each inhibitor was
strate the bandgap reduction by lanthanum doping (Fig. 7b, d, and added to the dye solution at a concentration of 1 mM, and
f). As we expected, under sunlight, the La/TiO2 and La/TiO2@HNTs photolysis was performed for 1 h under sunlight conditions. As
could achieve a higher decomposition yield and reaction rate than Fig. 8b shown, the degradation efficiency was 82.07% without
TiO2 and TiO2@HNTs, respectively. Compared with the previously inhibitors, in the presence of BQ, TEOA, and IPA was 49.95%,
reported catalysts with complicated fabrication, this synergistic 29.69%, and 70.40%, respectively (Fig. 8b). From this result, we
effect could render for the La/TiO2@HNTs to achieve the efficient could infer that the radical species were generated in the
 
decomposition of the dyes despite the simple synthetic method- quantitative order of h+ > O2 > OH during the photolysis process.
+
ology (Table S4). Recently, as an alternative photo-catalyst of TiO2, The generation of h in valance band by photons took a key role in
the ZnO nanoparticle were exploited, showing the 2.65 eV bandgap the dyes’ photolysis. The photolysis chemical reaction can be
[63]. Throught the alternation from TiO2 to ZnO might provide the expressed as follows:
similar enhancement of catalytic activity on sunlight. Additionally,
we performed the catalyst recycling experiment. After each (1) La/TiO2@HNTs + hv → e (CB) + h+ (VB)

photodegradation under UV light, the La/TiO2@HNTs catalysts (2) La/TiO2 (e) + O2 → O2

were separated by centrifugation, then dried and reused. The (3) TiO2@HNTs (h ) + H2O → OH + H+
+
+   
photolysis efficiency was maintained over 95% during five times (4) Dye + h + O2 + OH → CO2 + H2O

6
G Model
JIEC-5562; No. of Pages 8

J. Lee, S. Seong, S. Jin et al. Journal of Industrial and Engineering Chemistry xxx (xxxx) xxx–xxx

holes, decreasing the bandgap from 3.25 eV to 3.01 eV. This


combination of the HNT support and lanthanum doping on TiO2
showed synergistic photolysis activity enhancement, increasing
the catalytic activity by 3 times. The HNTs supports also offer an
easy separation method using centrifugation. The photoactive
range broadening induced by the lanthanum-doping allowed the
photolysis in sunlight without any extrinsic energy. The photo-
catalysts were systematically characterized by using XRD, Raman,
FT-IR, FE-SEM, TEM, and BET. The chemical mechanism of
photodegradation was investigated through the radical species
trapping, showing that hole generation is the key to photolysis. The
simple fabrication method and enhanced photolysis activity of our
photo-catalysts can be applied to purify an aquatic environment
contaminated with organic pollutants for the eco-friendly textile
industry. Biocompatibility studies and the natural degradation of
photo-catalysts would be required for use in industry. Thus, our
future work will be the development of degradable photo-
catalysts.

Conflicts of interest

None.

Acknowledgments

This work was supported by the Basic Science Research Program


through the National Research Foundation of Korea (NRF), funded
Fig. 7. Photos of photocatalytic degradation of rhodamine B with the catalysts
samples under (a) UV light and (b) direct sunlight. The photolysis of dyes with the
by the Ministry of Education (NRF-2012R1A6A1029029, NRF-
catalysts samples (c) under UV light and (d) sunlight in terms of time. Pseudo-first- 2018R1D1A1B07048063, NRF-2020R1A6A1A06046728). This re-
order kinetics curves of rhodamine B degradation reactions under (e) UV Lamp and search was also in part supported by the KIST Institutional Program
(f) direct sunlight. (2E31130, 2Z06270-20-138, 2V08610).

Appendix A. Supplementary data

Supplementary material related to this article can be


found, in the online version, at doi:https://doi.org/10.1016/j.
jiec.2021.05.029.

References

[1] D.T. Sponza, J. Hazard. Mater. 138 (2006) 438.


[2] S. Parvez, C. Venkataraman, S. Mukherji, Chemosphere 73 (2008) 1049.
[3] D.A. Yaseen, M. Scholz, Int. J. Environ. Sci. Technol. 16 (2019) 1193.
[4] A. Fujishima, K. Honda, Nature 238 (1972) 37.
[5] R. Asahi, T. Morikawa, T. Ohwaki, T. Aoki, K.Y. Taga, Science 293 (2001) 269.
[6] F. Tavakoli, M. Salavati-Niasari, A. badiei, F. Mohandes, Mater. Res. Bull. 63
Fig. 8. (a) Catalytically recyclable degradation of rhodamine B with La/TiO2@HNTs. (2015) 51.
(b) The photodegradation of rhodamine B after different scavengers were added [7] M. Salavati-Niasari, Z. Fereshteh, F. Davar, Polyhedron 28 (1) (2009) 126.
under direct sunlight. [8] M. Masjedi-Arani, M. Salavati-Niasari, Int. J. Hydrogen Energy 42 (27) (2017)
17184.
[9] F. Beshkar, H. Khojasteh, M. Salavati-Niasari, J. Colloid Interface Sci. 497
When TiO2 receives photons, electrons are excited from the (2017) 57.
balance band (VB) of TiO2 to the conduction band (CB), and h+ and [10] S. Zinatloo-Ajabshir, M. Salavati-Niasari, Z. Zinatloo-Ajabshir, Mater. Lett. 180
e– are generated (1). La ions on the surface of TiO2 trap excited (2016) 27.
[11] S. Zhu, D. Wang, Adv. Energy Mater. 7 (2017)1700841.
electrons, preventing recombination of electrons and hole pairs. [12] Y. Nosaka, A.Y. Nosaka, Chem. Rev. 117 (2017) 11302.

The generated holes (h+) and electrons (e) produced OH (3) and [13] K.C. Ko, S.T. Bromley, J.Y. Lee, F. Illas, J. Phys. Chem. Lett. 8 (2017) 5593.
  +   
O2 (4), respectively. The radical species h , O2 , and OH [14] F. Ansari, A. Sobhani, M. Salavati-Niasari, J. Colloid Interface Sci. 514 (2018) 723.
[15] A. Tolsosana-Moranchel, C. Pecharromán, M. Faraldos, A. Bahamonde, Chem.
completely decompose the dye into H2O, CO2, and byproducts (4). Eng. J. 403 (2021)126186.
[16] Subodh, K. Chaudhary, K. Prakash, D.T. Masram, Appl. Surf. Sci. 509 (2020)
Conclusions 144902.
[17] S. Mortazavi-Derazkola, M. Salavati-Niasari, O. Amiri, A. Abbasi, J. Energy
Chem. 26 (1) (2017) 17.
In conclusion, we have found the synergistic photolysis activity [18] B. Micó-Vicent, F.M. Martínez-Verdú, A. Novikov, A. Stavitskaya, V. Vinokurov,
of the La/TiO2@HNTs nanocomposite and their economical E. Rozhina, R. Fakhrullin, R. Yendluri, Y. Lvov, Adv. Funct. Mater. 28 (27) (2018)
1703553.
synthesis using sol–gel and sonochemical methods. As a support,
[19] S. Sadjadi, M. Malmir, G. Lazzara, G. Cavallaro, M.M. Heravi, Sci. Rep. 10 (1)
the HNTs improved the adsorption efficiency of substrates, (2020) 2039.
increasing the local concentration of substrate nearby the [20] G. Cavallaro, S. Milioto, G. Lazzara, Langmuir 36 (14) (2020) 3677.
photo-catalysts. The doping with lanthanum ions expanded the [21] S. Sadjadi, Appl. Clay Sci. 189 (2020)105537.
[22] S. Sadjadi, G. Lazzara, M. Malmir, M.M. Heravi, J. Catal. 366 (2018) 245.
photoactive range and prevented recombination of electrons and

7
G Model
JIEC-5562; No. of Pages 8

J. Lee, S. Seong, S. Jin et al. Journal of Industrial and Engineering Chemistry xxx (xxxx) xxx–xxx

[23] A.V. Stavitskaya, E.A. Kozlova, A.Y. Kurenkova, A.P. Glotov, D.S. Selischev, E.V. [46] T.S. Natarajan, K. Natarajan, H.C. Bajaj, R.J. Tayade, J. Nanopart. Res. 15 (2013)
Ivanov, D.V. Kozlov, V.A. Vinokurov, R.F. Fakhrullin, Y.M. Lvov, Chem. Eur. J. 26 1669.
(57) (2020)13085. [47] A.T.O. Dal Toe, G.L. Colpani, N. Padoin, M.A. Fiori, C. Soares, Appl. Surf. Sci. 441
[24] E. Rozhina, S. Batasheva, R. Miftakhova, X. Yan, A. Vikulina, D. Volodkin, R. (2018) 1057.
Fakhrullin, Appl. Clay Sci. 205 (2021)106041. [48] J. Guo, S. Zhu, Z. Chen, Y. Li, Z. Yu, Q. Liu, J. Li, C. Feng, D. Zhang, Ultrason.
[25] E. Tarasova, E. Naumenko, E. Rozhina, F. Akhatova, R. Fakhrullin, Appl. Clay Sci. Sonochem. 18 (2011) 1082.
169 (2019) 21. [49] P. Luo, B. Zhang, Y. Zhao, T. Wang, H. Zhang, J. Liu, Korean J. Chem. Eng. 28
[26] L. Ling, Y. Feng, H. Li, Y. Chen, J. Wen, J. Zhu, Z. Bian, Appl. Surf. Sci. 483 (2019) (2011) 800.
772. [50] Q.I. Rahman, M. Ahmad, S.K. Misra, M. Lohani, Mater. Lett. 91 (2013) 170.
[27] G. Mishra, M. Mukhopadhyay, Sci. Rep. 9 (2019) 40775. [51] J. Cui, D. Sun, W. Zhou, H. Liu, P. Hu, N. Ren, H. Qin, Z. Huang, J. Lin, H. Mac, Phys.
[28] A. Stavitskaya, K. Mazurova, M. Kotelev, O. Eliseev, P. Gushchin, A. Glotov, R. Chem. Chem. Phys. 13 (2011) 9232.
Kazantsev, V. Vinokurov, Y. Lvov, Molecules 25 (8) (2020) 1764. [52] N. Cai, Q. Dai, Z. Wang, X. Luo, Y. Xue, F. Yu, J. Mater. Sci. 50 (2015) 1435.
[29] Y. Lvov, A. Panchal, Y. Fu, R. Fakhrullin, M. Kryuchkova, S. Batasheva, A. [53] P. Sun, G. Liu, D. Ly, X. Dong, J. Wu, D. Wang, RSC Adv. 5 (2015) 52916.
Stavitskaya, A. Glotov, V. Vinokurov, Langmuir 35 (26) (2019) 8646. [54] Q. Zhang, L. Ma, M. Shao, J. Huang, M. Ding, X. Deng, X. Wei, X. Xu, J. Nanomater.
[30] N.S. Okten Besli, N. Orakdogen, Gels 7 (2021) 16. 2014 (2014)831752.
[31] T. Taroni, D. Meroni, K. Fidecka, D. Maggioni, M. Longhi, S. Ardizzone, Appl. [55] L. Lin, H. Wang, W. Jiang, A.R. Mkaouar, P. Xu, J. Hazard. Mater. 333 (2017) 162.
Surf. Sci. 486 (2019) 466. [56] B.H. Bac, N.T. Dung, L.Q. Khang, K.T. Hung, N.V. Lam, D.M. An, P.V. Son, T.T.V.
[32] X. Liu, J. Sun, S. Duan, Y. Wang, T. Hayat, A. Alsaedi, C. Wang, J. Li, Sci. Rep. 7 Anh, D.V. Chuong, B.T. Tinh, Minerals 8 (2018) 290.
(2017) 10033. [57] C. Li, Y. Zhao, T. Zhu, Y. Li, J. Ruan, G. Li, RSC Adv. 8 (2018) 14870.
[33] G. Lazzara, G. Cavallaro, A. Panchal, R. Fakhrullin, A. Stavitskaya, V. Vinokurov, [58] H.A. Mahmoud, K. Narasimharao, T.T. Ali, K.M.S. Khalil, Nanoscale Res. Lett. 13
Y. Lvov, Curr. Opin. Colloid Interface Sci. 35 (2018) 42. (2018) 48.
[34] H. Wang, D. Wu, X. Li, J. Mater. Sci.: Mater. Electron. 30 (2019) 19126. [59] S. Prathapani, V. More, S. Nohm, P. Bhargava, A. Yella, S. Mallick, Appl. Mater.
[35] B. Szczepanik, Appl. Clay Sci. 141 (2017) 277. Today 7 (2017) 112.
[36] S.U.M. Khan, M. Al-Shahry, W.B. Ingler Jr, Science 297 (2002) 2243. [60] J.F. Duarte Neto, I.D.S. Pereira, V.C. Da Silva, H.C. Ferreira, Gde A. Neves, R.R.
[37] S. Sakthivel, H. Kisch, Angew. Chem. Int. Ed. 42 (2003) 4908. Menezes, Ceramica 64 (2018) 598.
[38] S. Sakthivel, H. Kisch, ChemPhysChem 4 (2003) 487. [61] L. Li, H. Liu, W. Li, K. Liu, T. Tang, J. Liu, W. Jiang, Res. Chem. Intermed. 46 (2020)
[39] C. Burda, Y. Lou, X. Chen, A.C.S. Samia, J. Stout, J.L. Gole, Nano Lett. 3 (2003) 1049. 1715.
[40] T. Umebayashi, T. Yamaki, S. Tanaka, Nano Lett. 32 (2003) 330. [62] C.J. Stephenson, K.D. Shimizu, Org. Biomol. Chem. 8 (2010) 1027.
[41] H. Irie, S. Washizuka, N. Yoshino, K. Hashimoto, Chem. Commun. (2003) 1298. [63] M. Massaro, M. Casiello, L. D’Accolti, G. Lazzara, A. Nacci, G. Nicotra, R. Noto, A.
[42] H. Irie, Y. Watanabe, K. Hashimoto, J. Phys. Chem. B 107 (2003) 5483. Pettignano, C. Spinella, S. Riela, Appl. Clay Sci. 189 (2020)105527.
[43] J.L. Gole, J.D. Stout, C. Burda, Y. Lou, X. Chen, J. Phys. Chem. B 108 (2004) 1230. [64] E. Nyankson, N. Agyei-Tuffour, E. Annan, A. Yaya, B. Mensah, B. Onwona-
[44] T. Lindgren, J.M. Mwabora, E. Avendano, J. Jonsson, A. Hoel, C. Granqvist, S. Agyeman, R. Amedalor, B. Kwaku-Frimpong, J.K. Efavi, Heliyon 5 (2019)e01969.
Lindquist, J. Phys. Chem. B 107 (2003) 5709. [65] H. Wang, D. Wu, X. Li, P. Huo, J. Mater. Sci. Mater. Electron. 30 (2019) 19126.
[45] R. Jaiswal, J. Bharambe, N. Patel, Alpa Dashora, D.C. Kothari, A. Miotello, Appl.
Catal. B: Environ. 168–169 (2015) 333.

You might also like