You are on page 1of 22

Res Chem Intermed

DOI 10.1007/s11164-017-3067-4

Effect of hydrothermal conditions on microstructures


of pure and doped CeO2 nanoparticles and their photo-
catalytic activity: degradation mechanism and pathway
of methylene blue dye

Ahmed A. Farghali1 · Waleed M. A. El Rouby1 ·


Abdalrahman Hamdedein1

Received: 9 March 2017 / Accepted: 10 July 2017


© Springer Science+Business Media B.V. 2017

Abstract CeO2, Sm and Gd doped CeO2 were synthesized using the hydrothermal
route. The influence of synthesis conditions on the crystal structure, particle size and
microstructure was investigated. The prepared nanoparticles were characterized using
X-ray powder diffraction, FT-Raman spectroscopy, transmission electron microscopy
and scanning electron microscopy. The photo-catalytic degradation of methylene blue
was examined under UV light using the prepared nanoparticles at different conditions
(pH, dye concentration, catalyst dosage). The photo-catalytic activity of CeO2 was
kinetically enhanced by trivalent cations (Gd+3 and Sm+3) doping. The results displayed
the excellent catalytic activity of the Gd doped CeO2 as compared to that Sm doped
CeO2 and the undoped one. The degradation pathways of MB were followed using liquid
chromatography/mass spectroscopy (LC/MS), it was found that MB was degraded
completely into safe byproducts. The total organic carbon content measurements con-
firmed the results obtained by LC/MS.

& Ahmed A. Farghali


d_farghali@yahoo.com
1
Materials Science and Nanotechnology Department, Faculty of Postgraduate Studies for
Advanced Sciences (PSAS), Beni-Suef University, Beni-Suef, Egypt

123
A. A. Farghali et al.

Graphical Abstract

Keywords CeO2 · Doped CeO2 · Hydrothermal synthesis · Photo-catalytic


degradation · Pathway · Methylene blue

Introduction

Rare-earth oxides are extensively used in different applications, among them, pure
and doped cerium oxides (CeO2). CeO2 were employed as catalysts [1–3] cosmetic
materials [4, 5] ceramics [6], gas sensors [7–9] solid oxide fuel cells [10, 11] and
fluorescent materials [12, 13]. Accordingly, investigations on the properties of nano-
crystalline CeO2 have paid extensive interests during the past few years. Many
techniques such as sonochemical synthesis [14] hydrothermal synthesis processes
[15] gas condensation [16] and sol–gel processes [17] were used in the fabrication
of ceria.
Wastewater treatment is considered as a very important way to recycle water.
There are many ways to do it such as biological, coagulation, electrochemical
flotation, oxidation and adsorption. The latter is more desirable owing to large
uptake efficiency, ease of process and variety of adsorbents [18]. Novel
technologies with a higher efficiency and lower energy consumption are now
mandatory. An advanced oxidation process (AOPs), based on the generation of
reactive hydroxyl radicals (OH·) species that oxidize a broad range of organic
pollutants quickly and non-selectively is now under investigation.
Semiconductors are photo-catalysts classified as green technology used to
oxidize the organic pollutants into harmless inorganic materials like H2O and CO2
in the presence of light. In particular, nanosized pure and doped ceria have attracted
much attention with respect to the wastewater dye removal and the degradation of
other organic pollutants. Photo-catalytic oxidation using CeO2 is one of the
advanced oxidation processes being developed. When CeO2 is irradiated, the light
photons excite electrons to jump from the valence to the conduction band, leaving

123
Effect of hydrothermal condition on microstructures of…

holes in the former that will react with water molecules or hydroxide ions (OH) to
produce hydroxyl radicals (OH·) [19]. Oxygen is usually supplied as an electron
acceptor to prolong the recombination of electron–hole pairs during the photo-
catalytic oxidation. The oxidation of CeO2 is always carried out under UV light.
Accordingly, it is utilized in the photo-degradation of methylene blue that has wide
applications such as in coloring paper, temporary hair colorant, dyeing cottons and
wools. [20–22].
In this pieces of recent work, one highlighted the effect of the synthesis condition
on the physical characteristics of CeO2. The degradation of MB using CeO2 nano-
particles, either undoped or those doped, with one of the 4f elements (Gd+3/Sm+3)
was studied. The degradation pathway was followed for better understanding the
photo-catalytic degradation mechanism.

Experimental

Materials

Cerium nitrate (Ce(NO3), 99%), Ammonium ceric nitrate ((NH4)2Ce(NO3)6), 99%),


Gadolinium (III) acetate (Gd(Ac)3·xH2O), 99%), Samarium (III) nitrate (Sm
(NO3)·6H2O), 99%), and sodium hydroxide (NaOH, 98%) were purchased from(S d
fine—Chem limited—India) and used without further purification. Methylene blue
dye (99.7%) was purchased from (MERCK Millipore, Germany).

Samples preparation

In a typical experiment, 5.037 g of cerium nitrate (Ce(NO3)3·6H2O) was dissolved


in 100 mL distilled water. Then, NaOH (6 M) was added drop wise under vigorous
stirring for 30 min till complete precipitation of cerium hydroxide. The formed salt
solution was transferred into a Teflon lined stainless steel autoclave, and the
temperature was adjusted at different values (120, 160 and 200 °C) for different
times (12, 18 and 24 h). By the end of each experiment, the autoclave was allowed
to cool gradually at room temperature. Then, the precipitate was filtered, washed
several times, dried overnight and finally annealed at 600 °C in a muffle furnace for
3 h; the obtained material was donated as CeO2. 0.360 g of Samarium nitrate (Sm
(NO3)2·6H2O) or 0.366 g of Gadolinium acetate (Gd(CH3CO2)3xH2O) dissolved in
50 mL distilled water and mixed with a typical solution of 4.68 g cerium nitrate (Ce
(NO3)3·6H2O) dissolved in 50 ml distilled water in order to prepare 7% Sm or 7%
Gd doped CeO2 following the above mentioned conditions. For investigating the
precursor used and its effect on the properties of CeO2, ammonium ceric nitrate was
used instead of cerium nitrate under the same experimental conditions.

Powder analysis

X-ray diffraction experiments were conducted on a PANalytical (Empyrean) X-ray


diffractometer using Cu Kα radiation (wave length 0.154 cm−1) at an accelerating

123
A. A. Farghali et al.

voltage 40 kV, current of 35 mA, scan angle 20–70° range and scan step 0.02°. To
calculate the average crystallite size for the prepared nanoparticles, Scherrer’s
equation (Eq. 1) was used.
D ¼ 0:94k=b cos h; ð1Þ
where D is the average crystallite size, λ is the radiation wave length, β: the cor-
rected full width at half maximum of the diffraction peak and h is the diffraction
angle. FT-Raman spectra were recorded with a Bruker (Vertex 70 FTIR-FT Raman)
spectrometer with laser beam of 1064 nm power. Transmission Electron microscope
images were taken by JEOL-JEM 2100 (Japan) with an acceleration voltage of
200 kV. Field emission scanning electron micrographs photos were taken by Quanta
FEG 250 (Switzerland). The surface area measurements was performed using
Micrometrics Tri Star 2 for all prepared samples. The diffuse reflectance spectra of
all powder samples was recorded using a double beam UV–Vis-NIR spectropho-
tometer (Shimadzu UV-3600). LC/MS Quadrupole 6/20 (Agilent technologies
1260A225) was used to identify the intermediate products produced from the photo-
catalytic degradation of methylene blue dye.

Photo-catalytic activity

Photo-catalytic activity of pure CeO2 nanoparticles and their corresponding Sm and


Gd doped CeO2 nanoparticles were scrutinized for the degradation of methylene
blue dye (MB) under UV irradiation. Photo-catalytic experiments were performed at
298 K in a 150 ml beaker. Then, 0.1 gm of the prepared catalyst was added to
100 ml of 1.7 9 10−6 (M) MB solution, and then a UV light source with wavelength
254 nm was fixed vertically on the reaction at constant distance equal to 15 cm. At
specific time intervals, 1 ml of the sample solution was withdrawn from the reaction
medium and the change in the characteristic absorption at 662 cm−1 was followed
using an UV–Vis spectrophotometer (UV–Vis Jasco 530), from which the
concentration of MB was identified.
The rate of dye degradation was investigated by altering the pH values of the
starting solution, the dye concentration and the quantity of the catalyst used.

Results and discussion

X-ray diffraction studies

Effect of hydrothermal temperature

Figure 1a shows X-ray diffraction patterns of CeO2 nanoparticles prepared at


different hydrothermal temperatures (120, 160 and 200 °C, respectively) for 18 h.
The X-ray charts were compared with the ICDD no. (04-011-8929) assuming the
cubic symmetry of CeO2 nano-crystals. Strong reflections from (111), (200), (220),
(311), (222) and (400) planes of the cubic fluorite structure are seen in the patterns.

123
Effect of hydrothermal condition on microstructures of…

Fig. 1 XRD patterns of CeO2 nanoparticles prepared by hydrothermal method from cerium nitrate
precursor at: a different reaction temperatures (120, 160 and 200 °C for 18 h), b different irradiation times
(160 °C for 12, 18 and 24 h), c different precursors (cerium nitrate and ammonium ceric nitrate precursor
at 160 °C for 18 h), and d different doping materials (pure CeO2, Sm doped CeO2 and Gd doped CeO2
prepared at 160 °C for 18 h) [Ammonium cerium nitrate diffraction line]

This confirms the pure crystalline CeO2 nanoparticles. The appearance of a peak at
2-θ 44.5° as a secondary phase was indexed as ammonium cerium nitrate with
intensity depending on the hydrothermal temperature. This was interpreted due to
the use of ammonia in the adjustment of pH.
The lattice parameter was computed based on the cubic symmetry of the
crystalline lattice, and the theoretical density was calculated from Eq. 2.
Dx ¼ ZM=NV; ð2Þ
where Z is the number of molecules/unit cell, M is the sample molecule weight, V is
the unit cell volume (V = a3) and N is the Avogadros’number. The calculated
parameters are listed in Table 1, from which it was observed that the lattice
parameter was increased with increasing the hydrothermal temperature. It is not a
surprising result since the temperature increase gives more chance for lattice
expansion as well as the enhancement of crystal growth. This was reflected as a
monotonic decrease in the density since the unit cell was expanded. It was found
that as the temperature increased from 120, 160 to 200 °C, the average crystallite
size was increased to 13.8, 15.2 and 16 nm, respectively. Under hydrothermal

123
A. A. Farghali et al.

Table 1 Lattice parameters, theoretical density, crystallite size calculated from XRD, and particle size
calculated from TEM for all prepared CeO2 samples
Materials a (Å) Dx Crystallite size Particle size from
from XRD (nm) TEM (nm)

CeO2 nanoparticles prepared 5.4001 7.2589 13.8 10.7


at 120 °C for 18 h
CeO2 nanoparticles prepared 5.4146 7.2005 15.2 13.2
at 160 °C for 18 h
CeO2 nanoparticles prepared 5.4148 7.1999 16 10.9
at 200 °C for 18 h
CeO2 nanoparticles prepared 5.4139 7.2032 14 11.5
at 160 °C for 12 h
CeO2 nanoparticles prepared 5.4146 7.2005 15.2 13.2
at 160 °C for 18 h
CeO2 nanoparticles prepared 5.4188 7.1839 14.6 9.2
at 160 °C for 24 h
Sm doped CeO2 prepared at 5.40003 7.2589 15.7 8.5
160 °C for 18 h
Gd doped CeO2 prepared at 5.366 7.3979 15.6 10.4
160 °C for 18 h
CeO2 nanoparticles prepared 5.4146 7.2005 15.2 13.2
at 160 °C for 18 h from
cerium nitrate
CeO2 nanoparticles prepared 5.4588 7.0271 29.8 12.8
at 160 °C for 18 h from
amm. ceric nitrate

conditions, and during cerium hydroxide formation, the concentration of the


hydrogen ion (H+) is increased with increasing the reaction temperature, and this
high temperature in the hydrothermal reaction promotes the crystal growth in the
ceria particles from the hydrated cerium hydroxide complexes according to the
dissolution–precipitation mechanism [23].

Effect of reaction time

Samples prepared at different reaction times (12, 18 and 24 h) at 160 °C, are seen in
Fig. 1b. All reflections that appeared in the XRD patterns (Fig. 1b) were indexed
using ICDD card no. (04-011-8929). The former are indicating the face centered
cubic phase of CeO2 [24]. From crystallite size calculation it was found that, with
increasing the reaction time from for 12–18 h, the average crystallite size was
increased from 14 to 15.2 nm, respectively (Table 1). Increasing the time to 24 h,
the average crystallite size was slightly decreased to 14.65 nm. The crystallite size
increase may be attributed to the dissolution–precipitation mechanism by which the
crystal growth in the ceria particles is a function of the reaction duration. Moreover,
the average diffusion distance for the diffusing solute is short during cerium
hydroxide complex formation and many diffusing particles passing per unit time
and per unit area. Hence, the average dimension of crystallites depends on time

123
Effect of hydrothermal condition on microstructures of…

[23, 25]. The lattice parameter values increased and the density decreased (Table 1)
which means a thermal expansion owing to more reaction time, which has improved
the reaction rate hand in hand with the crystal growth.

Effect of precursor

CeO2 nanoparticles were synthesized from different precursors (cerium nitrate and
ammonium ceric nitrate) at 160 °C for 18 h and investigated using XRD. The same
diffraction patterns appeared corresponding to the cubic fluorite structure of CeO2.
It was found that CeO2 nanoparticles prepared using ammonium ceric nitrate
(Fig. 1c) reveal a crystallite size larger than that prepared from the cerium nitrate
precursor (Fig. 1c) at the same reaction conditions. The lattice parameter reflected
the same trend, while the density is always behaving in an opposite manner.

Effect of rare earth (Gd, Sm) doping on the CeO2

XRD patterns are shown in Fig. 1d for Sm and Gd doped CeO2, respectively. It was
noticed that the crystallite size changed from 15.2 to 15.7 and 12.6 nm in case of
pure CeO2, Sm doped CeO2 and Gd doped CeO2, respectively. Substitution of the
Ce4+ cations by a lower trivalent metal ion (e.g., Gd3+ or Sm3+) in the lattice results
in the oxygen vacancy formation and enlargement the crystal lattice. The lattice
expansion reduced the particle size and could be explained by increasing the point
defect concentrations as a result of substitution of Ce4+ by larger cations such as
Gd3+ [26, 27]. It is expected that the trivalent 4f ion (Gd+3 and/or Sm+3) enters the
cubic lattice at the expense of Ce+3 ions. Here, one could argue that both Gd+3 and
Sm+3 doping decreased the lattice parameter (Table 1). The direct reason is the
difference between the ionic radius of Ce+3 (1.143 Å) and that of Gd+3 (1.053 Å),
Sm+3 (1.079 Å) in 8f coordination. This decrease in the cubic lattice volume
resulted in an increase of density.

Raman spectroscopy analysis

Raman spectra of hydrothermally synthesized CeO2, Gd doped CeO2 and Sm doped


CeO2 powders are illustrated in Fig. 2, which confirm the formation of the fluorite
phase. The intensive band at 450–470 cm−1 corresponding to the allowed Raman
mode (F2g) of fluorite metal dioxides belonged to the O5h (Fm3 m) space group
[28–30]. For pure CeO2 powders, the Raman spectrum was symmetric around
464 cm−1, and the F2g mode corresponded to the symmetric vibration of oxygen ions
around Ce4+ ions [30]. In the Sm and Gd doped CeO2, the F2g band became
asymmetrical and slightly shifted to low frequencies, due to the difference between
Gd–O, Sm–O and Ce–O bond length as well as bond strength, that originated from
Gd+3 and/or Sm+3 induced changes in the cubic lattice of CeO2.

123
A. A. Farghali et al.

0.04 0.04

0.035
CeO
2 0.03
0.035 Sm doped CeO 0.025

Intensity
2
Gd doped CeO 0.02
2
0.03 0.015

0.01

0.005
0.025
0
Intensity

420 430 440 450 460 470 480 490 500


-1
Reman shift (Cm )

0.02

0.015

0.01

0.005

0
200 300 400 500 600 700 800
-1
Reman shift (Cm )

Fig. 2 Raman spectra of pure CeO2, Sm doped CeO2 and Gd doped CeO2 prepared at 160 °C for 18 h
hydrothermal time

Electron microscope examination

TEM analysis was used to scrutinize the size and shape of all the synthesized
nanoparticles. TEM images for CeO2 nanoparticles prepared at different hydrother-
mal temperatures (120, 160 and 200 °C) for 18 h are shown in Fig. 3a–c,
respectively. It is clear that CeO2 nanoparticles synthesized at 120 °C are closely
packed polygonal particles with slight agglomeration and particle size of 10.7 nm
(Fig. 3a; Table 1). While that synthesized at 160 °C have semispherical shape and
are agglomerated closely packed polygonal particles with increase in the average
particle size up to 13.2 nm (Fig. 3b; Table 1). With further increase in the
hydrothermal temperature value up to 200 °C, the resulting nanoparticles seem to be
more dispersed with average particle size 10.9 nm and no agglomeration was
observed (Fig. 3c; Table 1).
TEM micrographs of CeO2 nanoparticles synthesized at different reaction times
(12, 18 and 24 h) at 160 °C were used for evaluating the effect of hydrothermal
treatment time on the prepared CeO2 nanoparticles and are presented in Fig. 3d, b, e.
It is obviously clear that particles synthesized after 12 h have indefinite cubic shape
with agglomeration of the particles, and the average particle size was found to be
about 11.5 nm (Fig. 3d). By increasing the hydrothermal time to 18 h, the particles
are still agglomerated and the average particle size was increased to 13.2 nm
(Fig. 3b). With further increase in the hydrothermal time up to 24 h, it was noticed
that the edges of the polygonal shape of the particles tend to disappear to form semi-

123
Effect of hydrothermal condition on microstructures of…

Fig. 3 TEM images of CeO 2 nanoparticles prepared from cerium nitrate precursor at: a 120 °C for 18 h,
b 160 °C for 18 h, c 200 °C for 18 h, d 160 °C for 12 h, e 160 °C for 24 h, f 160 °C for 18 h using
ammonium ceric nitrate as a precursor, g Sm doped CeO2 prepared at 160 °C for 18 h and h Gd doped
CeO2 prepared at 160 °C for 18 h

spherical shapes with more dispersed particles and average particle size of 9.2 nm
(Fig. 3e).
Cerium oxide was prepared using two different precursors (ammonium ceric
nitrate and cCerium nitrate). TEM images illustrated that CeO2 samples obtained
from the cerium nitrate precursor has polygonal shapes and agglomerated particles
with particle size about 13.2 nm, while the other synthesized from ammonium ceric
nitrate has a more dispersed structure, and the particle size was found to be about
12.8 nm (Fig. 3f).
Figure 3g, h shows TEM images of Sm and Gd doped CeO2, respectively. It is
clearly observed that all samples had geometric shapes, although Gd doped CeO2
sample has more dispersal structure and nearly most particles have cubic symmetry
with clear shape boundaries with average particle size about 10.4 nm, while the Sm
doped CeO2 sample has average particle size about 8.52 nm. However, pure cerium
oxide has a larger particle size of about 13.2 nm with dense and agglomerated
structure as discussed previously.
It is clearly observed that there is a difference between the calculated crystallite
size from XRD and that observed from TEM analysis. For example, the average

123
A. A. Farghali et al.

particle size of Sm doped CeO2 is 15.7 nm from XRD analysis and 8.52 nm from
TEM. A certain density of dislocations inside the grains, which is thought to be
partly or fully responsible for the observed discrepancies, as indicated by TEM
imaging. The crystallite size was calculated from XRD using the Debye–Scherrer
equation, while that observed from HRTEM is the particle size the latter was
observed to have and measured on an average of ten particles, and while the former
was calculated from the FWHM of the most intense peak. The differences between
both values originated from microstrain, agglomeration of nanoparticles, electro-
static interaction and a magnetic character in some cases, see Table 1.
The surface morphology of the samples was investigated using SEM. Figure 4a–c
shows SEM images of pure CeO2, Sm doped CeO2 and Gd doped CeO2,
respectively. From Fig. 4a it is shown that highly condensed and agglomerated
particles were observed for pure CeO2 nanoparticles. While SEM images of Sm
doped CeO2 showed that small dense particles aggregate together to form larger
particles (Fig. 4b), whereas the Gd doped CeO2 image showed small and
homogeneous particles were obtained (Fig. 4c).

Surface area analysis

The nitrogen adsorption–desorption isotherms of all prepared samples are indicated


in Fig. 5a. According to the IUPAC classification, isotherms presented in Fig. 5a are
type IV isotherm with a H3 type hysteresis loop due to capillary condensation in
mesopores (2–50 nm) [31]. A typical feature of type IV isotherms is a final
saturation plateau of variable length, despite being sometimes reduced to an
inflection point. Hysteresis is a consequence of capillary condensation [31] (i.e., the
initial monolayer-multilayer adsorption on the walls of mesopores is followed by
pore condensation).
The pore size distribution curves were calculated from the desorption branch of
the isotherms by the BJH model as indicated in Fig. 5b. All samples have mono-
modal pore size distribution. All pore diameters are within the range of meso-pores
which is consistent with the existence of nanoparticles aggregates.
The BET specific surface areas of the prepared samples are given within Table 2.
The specific surface areas vary significantly from pure ceria to other doped samples.
Pure ceria provided the lowest surface area (31.96 m2 g−1) and Sm doped ceria
provided the surface area of (36.79 m2 g−1), while the highest specific surface area
obtained was that of Gd doped ceria (46.86 m2 g−1). The specific surface areas
corroborate well with the previously determined average crystallite sizes being
inversely proportional.

Optical properties

Optical band gap values, the nature of the transition and the effect of particle size all
have been determined. Figure 6 shows the optical band gap values of the prepared
samples depending on Tauc’s equation of direct allowed transition type [32–34].

123
Effect of hydrothermal condition on microstructures of…

Fig. 4 FESEM images of CeO2 nanoparticles prepared 160 °C for 18 h using cerium nitrate as a
precursor materials: a pure CeO2, b Sm doped CeO2 and c Gd doped CeO2

123
A. A. Farghali et al.

Fig. 5 a Nitrogen adsorption–desorption isotherms of the prepared samples. b BJH desorption pore
volume distribution of pure CeO2, Sm doped CeO2 and Gd doped CeO2 catalysts

Table 2 Surface area measurements for pure, Sm and Gd doped CeO2


Materials BET surface Average pore
area (m2 g−1) width (Å)

Pure CeO2 nanoparticles prepared at 160 °C for 18 h 31.96 84.02


from cerium nitrate
Sm doped CeO2 prepared at 160 °C for 18 h 36.79 110.23
Gd doped CeO2 prepared at 160 °C for 18 h 46.86 64.29

123
Effect of hydrothermal condition on microstructures of…

Fig. 6 The plots of (α hʋ)2 versus hʋ for pure, Sm and Gd-doped CeO2 prepared by hydrothermal
method

Table 3 Band gap energies of pure, Sm and Gd doped CeO2 nanoparticles


Materials Band gap (eV)

Pure CeO2 4.9


Sm doped CeO2 4.3
Gd doped CeO2 2.8

ðht  Eg Þ1=2
a¼A ð3Þ
ht
Table 3 shows the band gap values with the particle size obtained by different
means. As listed in the table, the recorded Eg value of pure CeO2 is 4.9 eV, which is
higher than the recorded value of bulk CeO2 (3.15 eV) [35, 36]. Such a result could
be attributed to the small particle size, whereas as the crystal size increased the band
gap energy decreased. Table 3 contains the band gap values of the entire samples.
Pure CeO2, Sm doped CeO2 and Gm- doped CeO2 exhibit a band gap values of 4.9,
4.3 and 2.8 eV, respectively, that are higher than that of the bulk value 3.15 eV, i.e.,
blue shifted.

Photo-catalytic activity measurements

Photo-catalytic degradation of MB dye under UV irradiation was studied as a model


reaction to investigate the photo-catalytic activity of the pure CeO2, Sm doped CeO2

123
A. A. Farghali et al.

and Gd doped CeO2 nanoparticles prepared by the hydrothermal method at 160 °C


for 18 h from cerium nitrate precursor. In the absence of photo-catalyst, the dye
solution subjected to UV irradiation was found to be stable.
At constant wavelength (662 cm−1), the change in the concentration of MB was
recorded following the UV irradiation time. The results obtained indicate that pure
nanocrystalline CeO2 catalyst under UV irradiation leads to complete decomposi-
tion of MB after 15 h, while Sm doped CeO2 nanoparticles can decompose MB
completely after 10 h. However, Gd doped CeO2 nanoparticles have the ability to
decompose MB completely after 9 h under UV light irradiation (See Fig. 7a).
Under UV radiation of CeO2, the photons excite the valence band electrons
across the band gap into the conduction band, leaving holes behind in the valence
band that will then react with water molecules or hydroxide ions (OH−) to produce
hydroxyl radicals (OH·) [37, 38]. These radicals are responsible for the degradation
process of MB. It is well known that doping increase the lattice free electrons and
holes; therefore, many hydroxyl ions will interact with the new formed holes,
increasing the rate of degradation of MB [39]. The rate of the photo-degradation of
MB not only depends on the lattice free radicals but also the morphology and the
crystallinity of the prepared samples. Nanocube-like morphology and crystallinity
which has been proven to be more favorable for the diffusion and separation of
photo-excited charge carriers affects greatly the rate of photo-degradation of MB.
So, Gd doped CeO2 has the best catalytic decomposition of MB due to its nano-cube
like morphology as indicated by TEM images (see Fig. 3h) [40].
It is well known that the primary reason for the enhanced activity of a catalyst is
its surface area value. Compared with pure ceria (31.96 m2 g−1) Sm doped ceria has
a higher surface area (36.79 m2 g−1) while Gd doped ceria has the highest surface
(46.86 m2 g−1) see Table 2. Moreover, the band gap values play a critical role in
determining the photo-catalytic activity of semiconductors. In our case Gd doped
CeO2 has the lowest bandgap value; thus, the energy required for exiting the valence
band electron will be low. Hence, Gd doped CeO2 has the best catalytic
decomposition of MB over pure and Sm doped ceria.
The kinetics of MB dye degradation in aqueous solution under UV light
irradiation was investigated, and the results are shown in Fig. 7b. It follows an
apparent first–order in agreement with a generally observed Langmuir–Hinshel-
wood kinetics model [41];
R ¼ dC=dt ¼ kKC =ð1 þ KC Þ; ð4Þ
where R is the degradation rate of the reactant (mg/min), C is the concentration of
the reactant (mg/l), t is the illumination time, k is the reaction rate constant (mg/
min) and K is the adsorption coefficient of the reactant (l/mg). This equation may be
simplified to;
LnðC0 =Ct Þ ¼ kKt ¼ kapp t: ð5Þ

By plotting Ln (C0/Ct) versus time (h), straight lines were obtained indicating
that the mechanism of all reactions obey first order kinetics.

123
Effect of hydrothermal condition on microstructures of…

Fig. 7 a Disappearance of MB by photocatalysis under UV-irradiation at a wavelength 662 cm−1 using


pure CeO2, Sm doped CeO2 and Gd doped CeO2 catalysts. b The pseudo-first-order degradation of MB
using pure CeO2, Sm doped CeO2 and Gd doped CeO2 catalysts

123
A. A. Farghali et al.

Effect of solution pH

The solution pH seems to play an important role in the photo-catalytic degradation


process. The effect of pH on the MB dye degradation was monitored in the range of
5–9. The pH value is adjusted to the desired value before the degradation process
using 0.1 (M) HCl and/or NaOH, and there is no pH control during the running
reaction. The rate of degradation at different pH solutions is shown in Fig. 8a. The
maximum degradation efficiency was obtained at a higher pH value of 9. Since the
pH of the solution greatly affected the surface charge properties and the size of
aggregates it forms on. The photo-catalytic process is generally generated from
electron–hole pairs formed on the catalyst’s surface irradiated by UV-light. The
hydroxyl radicals attract direct oxidation by positive holes and direct reduction by
the conducting band electrons. At higher pH solutions more and more hydroxyl ions
are obtained, which are ready to interact with the positive holes for more hydroxyl
radicals yield. Also, the negatively charged catalyst surface occurring in the basic
solution adsorb more positively charged MB dye molecules.

(a) 100 (b) 100

80 80
Degradation (%)
Degradation (%)

60 60

40 40
0.05 gm Catalyst

pH= 5 0.1 gm Catalyst


20 20
pH= 7 0.15 gm Catalyst
pH= 9

0 0
0 2 4 6 8 10 12 0 2 4 6 8 10 12
Time /h Time /h
(c) 100 (d) 120

100
80
Degradation (%)

Degradation (%)

80
60

60

40
40
-8
C = 1.7x10
MB
20 -7
C = 1.7x10
MB 20
-6
C = 1.7x10
MB

0 0
0 2 4 6 8 10 12 1 2 3 4 5
Time /h No of cycles

Fig. 8 a Effect of pH on photo-degradation of MB (20 mg/L) using Gd-doped CeO2 nanoparticles (1 g/


L). b Effect of initial MB dye concentrations on the catalytic degradation using Gd-doped CeO2
nanoparticles (1 g/L) at pH = 7. c Effect of catalyst dosage on the photo-catalytic degradation of MB
(20 mg/L) at pH = 7. d Effect of reusability of Gd-doped CeO2 on the photo-degradation of MB

123
Effect of hydrothermal condition on microstructures of…

Effect of initial MB concentration

Figure 8b, shows the effect of the initial concentration of MB (1.7 9 10−6,
1.7 9 10−7 and 1.7 9 10−8 M) on the rate of degradation using 0.1 g Gd-doped CeO2
as a catalyst. It was found that the rate of photo-degradation decreased with an
increase in the dye concentration. When the dye concentration increased, more dye
molecules adsorbed on the catalyst surface, which greatly affected the catalytic
behavior and blocked the active sites for the newer dye molecules. Also, the
increase in dye concentration reduced the path length of the photons entering the
dye solution. At higher dye concentration a significant amount of UV light may be
captured by the dye molecules instead of the catalyst surface, which reduce the
catalytic efficiency.

Effect of the amount of catalyst

The amount of the catalyst is a main parameter in the degradation process. It is


important to determine the appropriate quantity of the catalyst for efficient dye
removal in order to avoid the use of excess catalyst. Increasing the catalyst quantity
over the desired amount may cause a shielding effect and reduce the catalytic
activity. The degradation efficiency of 1.7 9 10−6 M of MB under UV irradiated at
different catalyst dosages 0.05, 0.1 and 0.15 g (Gd-doped CeO2) is illustrated in
Fig. 8c. It is clearly seen that the photo-degradation efficiency of MB increased at
higher photo-catalyst dosage of 0.15 g. This result may be attributed to the high
availability of active sites, which are ready to be attracted by the dye molecules.

The catalyst reusability

Gd-doped CeO2 efficiency was checked by means of accuracy and precision as


repeatability and reproducibility. It is well known that the reproducibility is one of
the most important parameters in photo-catalysis studies. In the present work five
replicate standard solutions of MB dye were tested under the same experimental
conditions (0.1 gm of catalyst for 9 h UV irradiation). After each cycle, the catalyst
was filtrated, washed with water and dried at 100 °C for 2 h. The cycling
performance of Gd-doped CeO2 for degrading MB solution is shown in Fig. 8d. It is
clearly observed that there is only a 14% reduction in degradation efficiency after
5-cycles. This result exhibits acceptable reproducibility.

Identification of photo-degradation intermediate products

The advanced oxidation process is one of the newest heterogeneous oxidation


methods, photo-catalysis appears as an emerging destructive technology leading to
the total mineralization of most of the organic pollutants [42–47]. Photo-catalytic
degradation of MB using Gd doped CeO2 as the best catalyst under UV irradiation
has been investigated in aqueous suspensions. In addition to the color removal of the
dye, Gd doped CeO2 was able to oxidize methylene blue dye, with a complete

123
A. A. Farghali et al.

mineralization of carbon, nitrogen and sulfur heteroatoms into CO2, NH4+, NO3−
and SO42−, respectively [48].
The intermediates generated during the degradation process were analyzed by
LC/MS at different time intervals and identified by comparison with commercial
standards by interpretation of their fragment ions in the mass spectra. Figure 9a–d
shows the mass spectra of MB dye and its degradation products at zero time (before
degradation), after 3, 6 and 9 h (final step), respectively. It has been attempted to
identify the main aromatic metabolites resulting from MB degradation process.
They are presented in Scheme 1, where they are logically reported according to their
decreasing molecular weight.
It is well know that the initiator of the degradation process are the two oxidative
agents: the photo-produced holes h+ (mainly involved in the decarboxylation
reaction (“photo-Kolbe) and/or the OH· radicals, which are known as strongly active

(a) 100 (b) 100

80 80
285
Intensity%(arb.units)
Intensity%(arb.units)

60 60

94
40 40
153

110 158
20 20 137 218
122 216
301
230
285

0 0
50 100 150 200 250 300 350 50 100 150 200 250 300 350

Mass(amu) Mass(amu)

(c) 15 (d) 50
218
40
Intensity%(arb.units)

Intensity%(arb.units)

156
10
30

137
285 20
5
110
301
10
94 230

0 0
50 100 150 200 250 300 350 50 100 150 200 250 300 350
Mass(amu) Mass(amu)

Fig. 9 Mass spectrum pathway obtained for the photo-catalytic degradation of MB under UV irradiation
in the presence of Gd doped CeO2 at different time intervals. a Before degradation, b after 3 h, c after 6 h
and d after 9 h

123
Effect of hydrothermal condition on microstructures of…

M= 284

N N
S

NH 2

M= 303
N N
S
O

NH 2
M= 215 NH 2
N SO 3 H
N

M= 136

NH 2 NH2 O
M= 230 H
HOC
N SO 3 H N OH N

M= 152 M= 137

NH 2
M= 218
H
N SO 3 H OH
O
H
N OH

M= 153 M= 94

OH
SO 3 H OH
OH
M= 158 M= 94
M=110

Scheme. 1 Photo-catalytic degradation pathway of methylene blue using doped CeO2

and degrading but non-selective agents and can be generated by the following
reactions [49]. The formed OH· radicals attack MB in the C–S+=C functional group.
Hence, the initial step of MB disintegration can be attributed to the cleavage of the
bonds of the C–S+=C functional group in MB (Eq. 4). Actually, the first product
obtained from the degradation of MB was the sulfoxide (identified at m/e = 303).
The sulfoxide group interacts with an OH· radical to produce the sulfone (non-
detected) and causing benzene ring opening (Eq. 6, 7). A sulfonic acids detected in
metabolites at m/e = 230, 218 and 158 are the result of a third OH· attack (Eq. 8).

123
A. A. Farghali et al.

NH2 C6 H3 ðRÞSð¼ OÞC6 H4 R þ OH ! NH2 C6 H3 ðRÞSO2 þ C6 H5 R


ð6Þ
and/or
NH2 C6 H3 ðRÞSð¼ OÞC6 H4 R þ OH ! NH2 C6 H4 R þ SO2 C6 H4 R
ð7Þ

SO2 C6 H4 R þ OH ! RC6 H4 SO3 H: ð8Þ


Two familiar and well-known pathways for the degradation of the three nitrogen
groups present in the MB structure are illustrated below: first of all the central
amino-group undergoes a N=C double bond cleavage. Secondly, The H· radicals
were used to saturate the two remained amino bonds, yielding a substituted aniline
as seen for products at m/e = 303, 230 and 218 in Scheme 1 [50–52]. The amino
group containing compounds and OH· radicals interact producing the corresponding
phenol compounds, estimated at m/e = 153 (Eq. 9).
RC6 H4 NH2 þ OH ! RC6 H4 OH þ NH2 ð9Þ

The aromatic ring will undergo hydroxylation producing phenolic metabolites,


with degradation that has already been studied. Before the ring opening the
hydroxyl hydroquinone has been the last aromatic compound detected. Many other
hydroxylated intermediates formed because of their poor extractability were
difficult to be detected owing to their hydrophilic character [53, 54].
Although the color removal of the dye is an important step in the degradation
process, this process may produce colorless and toxic byproducts, which add to the
cost of water treatment. It is well known that determining the degradation pathway
is very complicated and requires a large number of measurements. In order to prove
the continuous decrease in organic carbon with UV irradiation, the total organic
carbon (TOC) measurements were determined. The (TOC) elimination kinetics of
(MB) dye over Gd-doped CeO2 was determined. After nearly 7 h, the carbon
concentration was decreased by about 93% of the initial dye concentration. Indeed
no (MB) was left in the reaction medium, but after the degradation process, some
organic carbon was left (aldehydes and carboxylic acids) but no presence of any
aromatic rings. Moreover, the elimination rate becomes slow, and the reaction time
is longer for the complete degradation of (MB) dye. After 9 h irradiation, the
degradation efficiency was found to be 99.6% and the results are shown in Table 4.

Table 4 TOC measurements for MB solution before and after degradation using Gd doped CeO2
TOC Quantity/ppm Degradation %

Distilled water 2 –
Methylene blue 84 –
MB after photo-degradation using Gd-doped CeO2 2.3 99.64

123
Effect of hydrothermal condition on microstructures of…

Conclusion

In summary, nano-sized CeO2 and (sm/Gd) doped CeO2 were successfully


synthesized via the hydrothermal method. Different precursors were used for the
preparation of CeO2 nano-particles. Microstructure and particle shape analysis of
the prepared nanoparticles were analyzed using different techniques. The lattice
parameters and specific surface area were calculated for the prepared nanoparticles,
and it was found that it was affected by the preparation condition (Temperature,
time and precursor) and doping. The prepared nanoparticles were used as a photo-
catalyst for the degradation of MB dye under UV light. The obtained results
reflected the higher efficiency of the Gd doped CeO2 nanoparticles as a photo-
catalyst. It is able to oxidize methylene blue dye completely and has safe bi-
products as confirmed by GC–MS.

Acknowledgements This work was supported by the Science and Technology Development Fund
(STDF) in Egypt under the Egypt–US joint Project ID No. 4435. We are grateful to Associate Professor S.
I. El-Dek for her help in the lattice parameter discussion.

References
1. Z.R. Ismagilov, E.V. Matus, L.T. Tsikoza, Energy Environ. Sci. 1, 5 (2008)
2. D. Delimaris, T. Ioannides, Appl. Catal. B 89, 1 (2009)
3. W. El Rouby, A. Farghali, A. Hamdedein, Water Sci. Technol. 74, 10 (2016)
4. O.W.J. Kotov, P. Clements Isaac, Adv. Mater. 21, 3970–4004 (2009)
5. S. Yabe, T. Sato, J. Solid State Chem. 171, 7–11 (2003)
6. G. Hao, X. Liu, H. Wang, H. Be, L. Pei, W. Su, Solid State Ion. 225, 0 (2012)
7. N. Izu, W. Shin, N. Murayama, Sens. Actuators B Chem 93, 1–3 (2003)
8. Q. Cui, X. Dong, J. Wang, M. Li, J. Rare Earth 26, 5 (2008)
9. A.A. Aboud, H. Al-Kelesh, W.M. El Rouby, A.A. Farghali, A. Hamdedein, M.H. Khedr, J. Mater.
Res. Technol. (2017). doi:10.1016/j.jmrt.2017.03.003
10. Y. Zhai, S. Zhang, H. Pang, Mater. Lett. 61, 8–9 (2007)
11. M. Garcı́a-Melchor, N. López, J. Phys. Chem. C 118, 20 (2014)
12. M. Garcı́a-Melchor, N. López, J. Phys. Chem. C 118, 20 (2014)
13. K. Kaneko, K. Inoke, B. Freitag, A.B. Hungria, P.A. Midgley, T.W. Hansen, J. Zhang, S. Ohara, T.
Adschiri, Nano Lett. 7, 2 (2007)
14. D.V. Pinjari, A.B. Pandit, Ultrason. Sonochem. 18, 5 (2011)
15. L. Mädler, W.J. Stark, S.E. Pratsinis, J. Mater. Res. 17, 06 (2002)
16. X. Gao, Y. Jiang, Y. Zhong, Z. Luo, K. Cen, J. Hazard. Mater. 174, 1–3 (2010)
17. M.T. Uddin, M.A. Islam, S. Mahmud, M. Rukanuzzaman, J. Hazard. Mater. 164, 1 (2009)
18. M.I. Litter, M.E. Morgada, J. Bundschuh, Environ. Pollut. 158, 5 (2010)
19. S.B. Khan, M. Faisal, M.M. Rahman, A. Jamal, Sci. Total Environ. 409, 15 (2011)
20. C.A.P. Almeida, N.A. Debacher, A.J. Downs, L. Cottet, C.A.D. Mello, J. Colloid Interface Sci. 332, 1
(2009)
21. P. Waranusantigul, P. Pokethitiyook, M. Kruatrachue, E.S. Upatham, Environ. Pollut. 125, 3 (2003)
22. F.A. Pavan, A.C. Mazzocato, Y. Gushikem, Bioresour. Technol. 99, 8 (2008)
23. M.-H. Oh, J.-S. Nho, S.-B. Cho, J.-S. Lee, R.K. Singh, Mater. Chem. Phys. 124, 1 (2010)
24. C. Levy, C. Guizard, A. Julbe, J. Am. Ceram. Soc. 90, 3 (2007)
25. A. Bonamartini Corradi, F. Bondioli, A.M. Ferrari, T. Manfredini, Mater. Res. Bull. 41, 1 (2006)
26. S. Dikmen, J. Alloys Compd. 491, 1–2 (2010)
27. F. Zhang, S.-W. Chan, J.E. Spanier, E. Apak, Q. Jin, R.D. Robinson, I.P. Herman, Appl. Phys. Lett.
80, 1 (2002)
28. G. Li, R.L. Smith, H. Inomata, J. Am. Chem. Soc. 123, 44 (2001)
29. P. Singh, M.S. Hegde, Chem. Mater. 22, 3 (2009)

123
A. A. Farghali et al.

30. I. Kosacki, T. Suzuki, H.U. Anderson, P. Colomban, Solid State Ion. 149, 1–2 (2002)
31. M. Thommes, K. Kaneko, A.V. Neimark, J.P. Olivier, F. Rodriguez-Reinoso, J. Rouquerol, K.S.
Sing, Pure Appl. Chem. 87, 9–10 (2015)
32. H.-H. Ko, G. Yang, M.-C. Wang, X. Zhao, Ceram. Int. 40, 5 (2014)
33. S. Maensiri, C. Masingboon, P. Laokul, W. Jareonboon, V. Promarak, P.L. Anderson, S. Seraphin,
Cryst. Growth Des. 7, 5 (2007)
34. M. Taguchi, S. Takami, T. Adschiri, T. Nakane, K. Sato, T. Naka, CrystEngComm 13, 8 (2011)
35. S. Tsunekawa, T. Fukuda, A. Kasuya, J. Appl. Phys. 87, 3 (2000)
36. P. Patsalas, S. Logothetidis, C. Metaxa, Appl. Phys. Lett. 81, 3 (2002)
37. A. Houas, H. Lachheb, M. Ksibi, E. Elaloui, C. Guillard, J.-M. Herrmann, Appl. Catal. B Environ.
31, 2 (2001)
38. K. Tanaka, K. Padermpole, T. Hisanaga, Water Res. 34, 1 (2000)
39. G. Dutta, U.V. Waghmare, T. Baidya, M.S. Hegde, K.R. Priolkar, P.R. Sarode, Chem. Mater. 18, 14
(2006)
40. N. Sabari Arul, D. Mangalaraj, P.C. Chen, N. Ponpandian, C. Viswanathan, Mater. Lett. 65, 21–22
(2011)
41. A. Messerer, R. Niessner, U. Pöschl, Carbon 44, 2 (2006)
42. M. Pera-Titus, V. Garcı́a-Molina, M.A. Baños, J. Giménez, S. Esplugas, Appl. Catal. B Environ. 47,
4 (2004)
43. R. Andreozzi, V. Caprio, A. Insola, R. Marotta, Catal. Today 53, 1 (1999)
44. M. Klavarioti, D. Mantzavinos, D. Kassinos, Environ. Int. 35, 2 (2009)
45. D.F. Ollis, Cr Acad Sci Urss 3, 6 (2000)
46. I. Oller, S. Malato, J.A. Sánchez-Pérez, Sci. Total Environ. 409, 20 (2011)
47. X. Chen, L. Liu, P.Y. Yu, S.S. Mao, Science 331, 6018 (2011)
48. J.M. McCord, E.D. Day Jr., FEBS Lett. 86, 1 (1978)
49. K. Ikehata, M. Gamal El-Din, Ozone Sci. Eng. 27, 2 (2005)
50. S. Camiolo, P.A. Gale, M.B. Hursthouse, M.E. Light, Org. Biomol. Chem. 1, 4 (2003)
51. A. Khataee, M.B. Kasiri, J. Mol. Catal. A Chem. 328, 1 (2010)
52. C. Guillard, H. Lachheb, A. Houas, M. Ksibi, E. Elaloui, J.-M. Herrmann, J. Photochem. Photobiol.
A 158, 1 (2003)
53. G.K.C. Low, S.R. McEvoy, R.W. Matthews, Environ. Sci. Technol. 25, 3 (1991)
54. P. Calza, E. Pelizzetti, C. Minero, J. Appl. Electrochem. 35, 7–8 (2005)

123

You might also like