Вы находитесь на странице: 1из 10

Journal of Alloys and Compounds xxx (2015) xxxxxx

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Microwave-assisted hydrothermal synthesis of CePO4 nanostructures:


Correlation between the structural and optical properties
D. Palma-Ramrez a, M.A. Domnguez-Crespo a,, A.M. Torres-Huerta a, H. Dorantes-Rosales b,
E. Ramrez-Meneses c, E. Rodrguez a
a
b
c

Instituto Politcnico Nacional, CICATA-Unidad Altamira, Km 14.5, Carretera Tampico-Puerto Industrial Altamira, C.P. 89600 Altamira, Tamps, Mexico
Instituto Politcnico Nacional, ESIQIE, Departamento de Metalurgia, C.P. 07300 Mxico D.F., Mexico
Universidad Iberoamericana, Departamento de Ingeniera y Ciencias Qumicas, Prolongacin Paseo de la Reforma 880, Lomas de Santa Fe, C.P. 01219 Mxico D.F., Mexico

a r t i c l e

i n f o

Article history:
Available online xxxx
Keywords:
CePO4
Nanostructures
Microwave-assisted hydrothermal method

a b s t r a c t
In this work, the microwave-assisted hydrothermal method is proposed as an alternative to the synthesis
of cerium phosphate (CePO4) nanostructures to evaluate the inuence of different synthesis parameters
on both the structural and optical properties. In order to reach this goal, two different sets of experiments
were designed, varying the reaction temperature (130 and 180 C), synthesis time (15 and 30 min) and
sintering temperature (400 and 600 C), maintaining a constant pH = 3. Thereafter, two experimental
conditions were selected to assess changes in the properties of CePO4 nanopowders with pH (1, 5, 9
and 11). The crystal structure and morphology of the nanostructures were characterized by X-ray diffraction (XRD), transmission electron microscopy (TEM) and scanning electron microscopy (SEM), respectively. Diffuse reectance properties of CePO4 with different microstructures were studied. The results
demonstrated that by using the microwave-assisted hydrothermal method, the shape, size and structural
phase of CePO4 can be modulated by using relatively low synthesis temperatures and short reaction
times, and depending on pH, a sintering process is not needed to obtain either a desired phase or size.
Under the selected experimental conditions, the materials underwent an evolution from nanorods to
semispherical nanoparticles, accompanied by a phase transition from hexagonal to monoclinic.
2014 Elsevier B.V. All rights reserved.

1. Introduction
In recent years, CePO4 nanostructures have become increasingly
important in a variety of applications such as uorescence, ion
exchange, catalytic materials and ceramic composite materials
with high mechanical properties [1] as a result of their low dimensionality and the quantum connement effect. The technological
application of these nanostructured materials is strongly dependent on their morphology, crystalline phase and particle size.
CePO4 presents two phases: monoclinic and hexagonal [2,3]. Few
studies on the production of CePO4 nanostructured materials have
been found in the literature. The hexagonal phase can be easily
obtained at low temperatures, while the monoclinic phase could
be prepared via the solid state reaction and hydrothermal method
at high temperatures [2]. Among these methods, the hydrothermal
synthesis seems to be potentially useful to obtain CePO4 nanostructures; the preparation of inorganic nano or micromaterials
Corresponding author.
E-mail address: mdominguezc@ipn.mx (M.A. Domnguez-Crespo).

by this technique features advantages such as the use of simple


equipment, low cost, high-uniform area production, low process
temperatures, catalyst-free growth, environmental friendliness
and very-easy-to-control particle sizes; however, it usually
requires long reaction times (about 24 h) to reach the nanoscale
[3,4]. As an enhancement of the typical hydrothermal method,
the microwave-assisted route has the advantage of producing
nanostructures with shorter synthesis times as a consequence of
the efcient and fast heating during the reaction process [5]. Particularly for CePO4, Ekthammathat et al. [6] have reported that
monoclinic nanorods can be synthesized from Ce(NO3)36H2O
and Na3PO412H2O at pH 1 by a simple microwave radiation for
60 min, while Patra and coworkers [7] obtained rhabdophane-type
hexagonal nanorods by the microwave-assisted solvothermal
synthesis, using a domestic microwave oven.
To the best of our knowledge, there have been few published
studies related to the synthesis of CePO4 nanostructures using
microwave energy. The aim of this work is to determine the effects
of the synthesis parameters on the structural, morphological
and optical properties of CePO4 nanostructures produced by the

http://dx.doi.org/10.1016/j.jallcom.2014.12.053
0925-8388/ 2014 Elsevier B.V. All rights reserved.

Please cite this article in press as: D. Palma-Ramrez et al., J. Alloys Comp. (2015), http://dx.doi.org/10.1016/j.jallcom.2014.12.053

D. Palma-Ramrez et al. / Journal of Alloys and Compounds xxx (2015) xxxxxx


were added dropwise to 25 ml of H5P3O10 under stirring; then, deionized water
was added to adjust a nal volume of 100 ml at pH 3. Thereafter, solutions were
transferred into a Teon container (autoclave). The autoclave was sealed and heated
for each experiment by varying both the synthesis temperature (130 and 180 C)
and time (15 and 30 min). Subsequently, the precipitate was separated by ltration
and dried for 24 h at 90 C and sintered at two different temperatures (see Table 1).
From these initial experiments, the synthesis conditions of the samples labeled
as 7 and 8 were chosen to evaluate changes in the morphology and particle size of
CePO4 nanopowders at different pH values. These effects were analyzed at pH = 1, 5,
9, and 11. HNO3 (37%) and NH4OH (30%) were used to adjust the pH values either in
alkaline or acid medium.

Table 1
Experimental design used to obtain CePO4 nanoparticles using a pH 3.
Experiment
at pH 3

Reaction
temperature
(Tr, C)

Synthesis
time
(t, min)

Sintering
temperature
(Ts, C)

1
2
3
4
5
6
7
8

130
180
130
180
130
180
130
180

15
15
30
30
15
15
30
30

400
400
400
400
600
600
600
600

2.2. Characterization of nanopowders

microwave-assisted hydrothermal method. For this purpose, variables such as reaction time, synthesis and sintering temperature
as well as pH of the media have been evaluated.

2. Experimental details
2.1. Preparation of CePO4 nanostructures
Tripolyphosphoric acid solution (0.035 mol1, H5P3O10) was obtained by using a
cation exchange resin (Dowex 50W X4 100-200 mesh) for conversion of sodium tripolyphosphate, Na5P3O105H2O (purum p.a., P98.0% (T), SigmaAldrich).
CePO4 nanostructures were synthesized by mixing Ce(NO3)36H2O (0.10 mol)
(SigmaAldrich, 99% trace metals basis) and H5P3O10 (0.10 mol). The microwave
reaction was performed in a microwave oven (CEM-MARS, frequency 2.45 GHz,
power of 200 W). The synthesis process was as follows: 50 ml of Ce(NO3)36H2O

The structure of the CePO4 nanoparticles was determined by X-ray powder diffraction (XRD) with a Bruker D8 Advance diffractometer equipped with a Lynxeye
detector and Cu Ka radiation (k = 1.5405 ) at 35 kV and 25 mA. Data were collected
at room temperature in the 2h range of 1570, step size of 0.016 and step time of
0.5 s. To evaluate the effects of the reaction parameters and sintering temperature
on the structural and vibrational properties of CePO4, Fourier transform infrared
spectroscopy (FTIR) was carried out. The spectra were recorded on a Perkin Elmer
spectrometer using KBr pellets, 4 cm1 of resolution setting and range of 1300
450 cm1. The samples were scanned 40 times. Scanning electron microscopy
was used to evaluate morphological changes in the CePO4 nanopowders using a
JEOL JSM-6300 apparatus (20 kV), whereas transmission electron microscopy was
used to corroborate the structure and phase composition of the samples by using
a JEOL-2000 FX-II working at an accelerating voltage of 100 kV. The analysis of
the average particle size and size distribution were determined by dynamic light
scattering, using deionized water as dispersant medium in a Malver Zetasizer Nano
ZSP, model ZEN5600. The particle size of the dispersions was described by the
cumulants mean diameter, and the size distribution was described by the polydispersity index and the size distribution plot. The optical properties of the samples
were followed by ultravioletvisible diffuse reectance spectroscopy (UVVis DR)
using a 110-mm-diameter-integrating-sphere accessory mounted on a Cary 5000

(b)

(a)

Monoclinic Hexagonal

Tr180t30Ts600

Tr180Ct30min

Tr130t30Ts600

(321)

30

40

50

60

70

20

30

50

60

Tr130t30Ts600pH11

(d)

Tr180t30Ts600pH11

Tr130 t30Ts0pH11

Tr180t30Ts0pH11

60

20

30

40

(033)

(402)

(233)
(041)

(212)
(320)
(023)
(040)

(031)
(221)

Tr180t30Ts600pH1
(012)
(202)
(212)

(101)

(101)
(011)

(402)
(321)

70

(311)

(121)

Hexagonal
Monoclinic
(020)
(200)
(120)

(142)

(231)
(132)

(033)

(321)
(103)

50

2 (degrees)

Tr180t30Ts600pH5

Tr180t30Ts0pH5

Tr130t30Ts0pH1

(003)
(301)
(212)
(113)
(220)
(302)
(310)
(311)
(213)
(104)
(222)
(312)
(041)
(114)

(112)
(210)
(022)
(211)

(111)

(200)
(120)
(012)
(012)

(110)
(200)

(202)
(212)
(210)
(022)
(031)
(311)

40

Tr130t30Ts600 pH1

Tr180t30Ts0pH9

(211)

(112)

Tr130t30Ts600pH5

Intensity (a.u.)

Tr130t30Ts0pH9

(214)

(402)

(121)

Tr180t30Ts600pH9

Tr130t30Ts0pH5

(101)

Intensity (a.u.)

Tr130t30Ts600pH9

30

70

2 (degrees)

(c)

20

Tr130t15Ts400
(321)

(103)
(212)
(200)
(302)
(310)
(104)

40

2 (degrees)

Hexagonal
Monoclinic

(004)
(142)
(241)

(031)
(311)

(212)
(103)
(023)
(132)
(140)

(020)
(200)
(120)
(012)
(202)
(212)

(011)
(111)

Tr130t30Ts400
Tr180t15Ts400

Tr130Ct15min
20

Tr130t15Ts600

Tr180t30Ts400

(101)

(303)

(104)

(301)
(212)
(220)
(302)
(310)

(112)

(211)
(003)

(110)

(200)
(102)

Rhabdophane (hexagonal)

(112)
(211)

Tr130Ct 30min

Intensity (a.u.)

Tr180Ct15min

(101)

Intensity (a.u.)

Tr180t15Ts600

50

Tr180t30Ts0pH1

60

70

2 (degrees)

Fig. 1. X-ray diffraction patterns of (a) non-sintered CePO4 powders obtained under the stated conditions, (b) CePO4 nanoparticles sintered at two different temperatures
(400 and 600 C), (c) samples synthesized under experiment 7 conditions at different solution pH values and (d) samples synthesized under experiment 8 conditions at
different solution pH values.

Please cite this article in press as: D. Palma-Ramrez et al., J. Alloys Comp. (2015), http://dx.doi.org/10.1016/j.jallcom.2014.12.053

D. Palma-Ramrez et al. / Journal of Alloys and Compounds xxx (2015) xxxxxx


Table 2
Crystal size of CePO4 nanostructures estimated from Scherrer equation.
Experiment

Crystal size of powders


obtained at pH 3
Ts0 (nm)

Ts400

1
2
3
4
5
6
7

16
14
17
16
16
14
17

15
15
15
17
19
21
20

16

20

or 600C

Crystal size of powders


obtained at pH 1, 5, 9 and 11
(nm)

Ts0 (nm)

Ts400

52 (pH 1)
11 (pH 5)
15 (pH 9)
16 (pH 11)
29 (pH 1)
8 (pH 5)
13 (pH 9)
10 (pH 11)

15
11
14
17
18
12
14
10

or 600C

(pH
(pH
(pH
(pH
(pH
(pH
(pH
(pH

Table 4
Particle z-average diameter (Dz), polydispersity index (PDI), the highest mean number
(%) and their sizes of CePO4 nanopowders.
Condition

Dz (nm)

PDI

The highest
mean number (%)

Size (nm)

1
2
3
4
5
6
7

529.2
541.3
445.3
533.4
315
396
pH 1 (335.4)
pH 3 (497.7)
pH 5 (270.8)
pH 9 (228.1)
pH 11 (214.4)
pH 1 (370.7)
pH 3 (513.9)
pH 5 (295.7)
pH 9 (266)
pH 11 (201.3)

1.16
1.1
1.12
1.13
1.1
1.11
pH 1 (1.82)
pH 3 (1.16)
pH 5 (1.39)
pH 9 (1.31)
pH 11 (1.48)
pH 1 (1.35)
pH 3 (1.14)
pH 5 (1.19)
pH 9 (1.06)
pH 11 (1.27)

26
31
28
24
31
27
14
24
18
21
21
19
24
27
56
22

396
459
342
459
255
342
122 (pH 1)
396 (pH 3)
141 (pH 5)
122 (pH 9)
91 (pH 11)
190 (pH 1)
255 (pH 3)
190 (pH 5)
14 (pH 9)
122 (pH 11)

(nm)

1)
5)
9)
11)
1)
5)
9)
11)

spectrophotometer. Samples were scanned with an average time of 0.1 s, scan rate
of 600 nm/min and data intervals of 1 nm within the 700200 nm range. The background reectance of polytetrauoroethylene (PTFE) was measured before. The
Kubelka Munk function F(R1) was applied to convert the diffuse reectance spectra
into the equivalent absorption spectra and determine the optical band gap from the
slope of the linear part in the (F(R1)hv)2 vs. hv plot (photon energy).

3. Results and discussion


3.1. Structure and morphology of the samples
Fig. 1a and b shows the results obtained from the X-ray diffraction analysis of CePO4 nanoparticles before and after being subjected to sintering in the rst experiment series. As it can be
seen in Fig. 1a, the peaks of the CePO4 powders match with the
hexagonal phase, known as rhabdophane (PDF card # 04-0125051). The main differences between the synthesis temperatures
in the microwave process are observed in the (1 0 1) reection,
which disappears at 180 C. A strong relationship between the
structure type and temperature has been reported in the literature
for CePO4 materials [4,8,9], where the rhabdophane structure
begins to be transformed into the monoclinic phase by heating
the nanopowders at temperatures above 400 C. Regularly, an
anhydrous form of the hexagonal phase is obtained between 100
and 400 C; however, depending on the cation used during the synthesis, a transition of monoclinic monazite can be completely
reached at 650 C [10]. In our case, the hexagonal phase tends to

(pH
(pH
(pH
(pH
(pH
(pH
(pH
(pH
(pH
(pH

1)
3)
5)
9)
11)
1)
3)
5)
9)
11)

grow with preferential direction at 180 C without the monazite


formation.
To evaluate possible effects of the synthesis temperature of the
CePO4 nanostructures on the hexagonalmonoclinic phase transformation, the samples were sintered at two different temperatures (400 and 600 C) and the results are presented in Fig. 1b.
The XRD patterns indicate that at 400 C, both structural phases
(hexagonal and monoclinic) coexist, while at 600 C, the peaks
agree with the monoclinic structure of the CePO4 powders (PDF

card # 04-007-2786). The (2 1 1) and (311)
planes, corresponding
to the hexagonal and monoclinic structures, respectively, overlapped each other at 400 C, but a combination of the synthesis
and sintering temperatures split the peak ca. 42, forming a new
(0 3 1) plane. Due to the fact that at 400 C only the hexagonal
phase is reported, the existence of both phases can be strongly
inuenced by the microwave supplied energy. Furthermore, the
phase transformation between the rhabdophane-type structure
(hexagonal system) and the monazite-type structure (monoclinic
system) has been observed from about 650 C [11], but in this case,
the monoclinic structure can be completely achieved at low sintering temperatures, highlighting the effect of the microwave energy.
The interesting point here is that by using microwave energy,
there is a great inuence on the microstructure and as a consequence, nanostructures with specic properties can be evaluated.
To evaluate the dependence of the nanopowder structure on pH,
a second set of experiments was designed by adjusting the pH

Table 3
Crystal size of CePO4 and treatment of linear plots to obtain the size of crystallites by Scherrer modied formula.
Condition

As-prepared eln

1
2
3
4
5
6
7
8
7
7
7
7
8
8
8
8

e4:4404 0.0117
e4:5628 0.0104
e4:6342 0.0097
e4:58 0.0102
e4:4404 0.0117
e4:5628 0.0104
e4:6342 0.0097
e4:58 0.0102
e5:955 0.0025
e4:5992 0.0010
e4:8082 0.0081
e4:8469 0.0078
e5:70206 0.0033
e4:5758 0.0102
e4:70319 0.0090
e4:51318 0.0109

at pH 3
at pH 3
at pH 3
at pH 3
at pH 3
at pH 3
at pH 3
at pH 3
(pH 1)
(pH 5)
(pH 9)
(pH 11)
(pH 1)
(pH 5)
(pH 9)
(pH 11)

Kk=L

0:94x0:154051
L

L (nm)

Sintered eln

12
14
15
14
12
14
15
14
56
14
18
18
43
14
16
13

e4:581 0.0102
e4:5222 0.0108
e4:63076 0.0097
e4:6 0.0100
e4:825 0.0080
e4:938 0.0071
e4:851 0.0078
e4:824 0.0080
e4:22391 0.00146
e2:9174 0.0540
e4:7569 0.0085
e4:9299 0.0072
e4:6425 0.0096
e2:9742 0.0510
e4:9476 0.0071
e4:5768 0.0102

Kk=L

0:94x0:154051
L

L (nm)
14
13
15
14
18
20
19
18
10
3
17
20
15
3
20
14

Please cite this article in press as: D. Palma-Ramrez et al., J. Alloys Comp. (2015), http://dx.doi.org/10.1016/j.jallcom.2014.12.053

D. Palma-Ramrez et al. / Journal of Alloys and Compounds xxx (2015) xxxxxx

(a)

( 1) h (100)
(2) m (101)

50 nm

(3) m (020)
(4) m (002)
(5) h (012)

20 nm

(6) m (

( 1) h (100)

(b)

(2) h (101)

100nm

(3) m (

(4) m (

(5) h (112)
(6) h (

20 nm

(7) m (310)
(1) m (011)

(c)

(2) m (020)

100nm

(3) m (

(4) m (

(5) m (013)

20 nm

(6) m (

(7) m (

(1) m (201)

(d)

(2) m (002)

100nm

(3) m (

(4) m (131)
(5) m (301)
(6) m (310)

50 nm

(7) m (

Fig. 2. Selected TEM micrographs and SAED patterns of CePO4 nanopowders synthesized under different experimental conditions: (a) Tr130t15Ts400, (b) Tr130t30Ts400, (c)
Tr180t15Ts600 and (d) Tr180t30Ts600. In the gure, m and h represent the monoclinic and hexagonal phases, respectively.

solution at 1, 5, 9, and 11, using the synthesis conditions of samples


7 and 8. Additionally, to highlight the structural modication with
the sintering temperature, XRD patterns of these samples were
scanned at room temperature and were labeled as Ts0.
Fig. 1c and d shows the XRD patterns of CePO4 synthesized at
solution pH values of 1, 5, 9 and 11 (Tr = 130 C). The XRD results
show some structural changes with pH even before starting the
sintering process. For example, for the samples synthesized under
the conditions of experiment seven (Fig. 1c), it is seen that by using
a solution with pH = 1, the hexagonal phase is favored, whereas a
transformation of this phase into the monoclinic one took place
at pH = 5, where both structural phases (hexagonal and monoclinic) coexist; such transformation seems to be reached in alkaline
medium (pH = 9 and 11). It is important to mention that there is a
discrepancy with a previous work using similar experimental conditions, where the monoclinic structure was observed at pH 1,

whereas the hexagonal phase coexists by adjusting the solution


in the pH range of 25 [6]. The mismatch in the results is mainly
correlated with the type of precursor used in the synthesis
(Na3PO412H2O or H5P3O10) and its interaction with Ce(NO3)3
during the synthesis. Even, the authors obtained similar results
with other rare earth elements to obtain LaPO4 nanorods [12]. This
observation pointed out the inuence of the acid or alkaline
character of the raw materials to obtain a desired structure. As
expected, after the thermal treatment at 600 C, the samples
crystallized into the monazite form.
On the other hand, by increasing the synthesis temperature up
to Tr = 180 C for 30 min, the monoclinic structure is favored, even
after the treatment at 600 C (Fig. 1d). In this case, the XRD patterns evaluated before the sintering process displayed less reections of the hexagonal structure at pH = 1, indicating that at this
reaction temperature (180 C), the transformation from hexagonal

Please cite this article in press as: D. Palma-Ramrez et al., J. Alloys Comp. (2015), http://dx.doi.org/10.1016/j.jallcom.2014.12.053

D. Palma-Ramrez et al. / Journal of Alloys and Compounds xxx (2015) xxxxxx

Experiment 7 (Tr130,t30,Ts600)

(a)

(c)

(b)

pH= 1

100 nm

pH= 5

100 nm

(d)

pH= 9

100 nm

pH= 11

100 nm

Experiment 8 (Tr180,t30,Ts600)

(e)

(f)

pH= 1

100 nm

pH= 5

(h)

(g)

100 nm

pH= 9

100 nm

pH= 11

Fig. 3. SEM micrographs of CePO4 nanopowders obtained under the conditions of experiments: (ad) 7 and (eh) 8.

Table 5
Band gap of CePO4 nanopowders calculated from the KubelkaMunk modied
spectra.
Experiment (pH 3)

Band gap (eV)

1
2
3
4
5
6
7

3.0
3.07
3.0
3.07
3.06
3.1
3.02

3.07

pH
pH
pH
pH
pH
pH
pH
pH

Kk
Kk 1

L  cos h
L cos h
Kk
Kk
1
ln b ln
ln
ln
L  cos h
L
cos h

1
5
9
11
1
5
9
11

3.22
3.04
3.25
3.25
3.15
3.25
3.08
3.23

to monoclinic has begun. Evidently, a pure monazite structure


without impurities of the rhabdophane phase was obtained after
sintering the samples at 600 C. The monazite structure is clearly
obtained at pH values of 5, 9 or 11 without the annealing process
and the polycrystallinity of the phase is enhanced at 600 C.
A quantitative estimation of the domain size for CePO4 nanopowders was evaluated and shown in Table 2 from Scherer equation, using the most intense peak for each XRD pattern:

Kk
b cos h

obtain the average L value through all the peaks (or some selected
peaks), the mathematical errors in the calculation must be reduced
by using the least square method to obtain the following equations:

where L = nanocrystallite size, k = wavelength of X-ray radiation (Cu


Ka, k = 0.15405 nm), b = FWHM, full width at half maximum of
peaks (radians), h = Bragg angle, K = is the shape factor, which can
vary from 0.62 to 2.08, where 0.89 is usually taken for spherical
crystals [13]. Thus, K depends on at least three things: breadth definition, crystallite shape and crystallite-size distribution [14]. In the
analyses, the factor value (K) was considered for a spherical shape
(0.89). Additionally, Scherrer results were compared with the
results obtained by a Scherrer modied method developed by Monshi et al. [15]. The modication used in this work correct the systematic error stemming from the assumption that there are N
different peaks of specic nanocrystals at either the 0180 (2h)
or 090 (h) range; then, all these N peaks must present identical
L values for the crystal size, which is not necessary true. Thus, to

2
3

After applying the least square method, the slope ln Kk


can be
L
obtained and the plot intercept can be calculated using the following expression [15].
Kk

eln L

Kk
L

For the calculations, the K value was considered to vary from 0.90,
for semispherical nanoparticles, to 0.94 for nanorod [16] structure
and the results from both set of experiments have been summarized in Tables 2 and 3.
The rst set of experiments displays crystal sizes around 14
21 nm, which indicates that the sintering temperature exerts no
great inuence on the crystallite size, while a similar trend was
obtained during the second stage. Under the conditions of experiments 7 or 8 at pH values of 5, 9 and 11, the CePO4 nanopowders
present values in the 816 nm and 1017 nm ranges at room temperature and 600 C, respectively. On the other hand, the CePO4
nanopowders obtained at pH = 1 in both experiments displayed
the largest crystallite sizes (52 and 29 nm) with an important
reduction after the sintering process (29 and 18 nm). In the analyses, it was assumed that the particle shapes were spherical, however, it has been reported that CePO4 can have either a rod or
spherical morphology [8]. Thus, the crystal size change after the
sintering process could be due to the transformation from rods into
spherical particles, which could be related to the smaller amount of
rods present in the sample, and therefore, the broadening of the
diffraction peaks due to the formation of smaller clusters is
observed [17].
As it can be seen in Table 3, strong associations were found
between the crystal size obtained by Scherrer and Scherrer modied equations. The rst and second sets of experiments reveal
crystal sizes between 1220 nm and 1318 nm, respectively. By
comparing the results obtained from Scherrer equation, it is conrmed that the behavior of the crystal size shows no big differences among the reaction temperature, sintering temperature

Please cite this article in press as: D. Palma-Ramrez et al., J. Alloys Comp. (2015), http://dx.doi.org/10.1016/j.jallcom.2014.12.053

D. Palma-Ramrez et al. / Journal of Alloys and Compounds xxx (2015) xxxxxx

(a)

(b)

pH 11

20
Tr180t15Ts600

20

Number (%)

Number (%)

30

25

Tr130t30Ts400

Tr130t30Ts600
Tr180t30Ts400

Tr130t15Ts600

Tr180t15Ts400

10

pH 9

Tr130t30Ts600

pH 5

15
10

pH 1

5
Tr130t15Ts400

0
0

100 200 300 400 500 600 700 800 900

100

Particle size (nm)


40

200

300

400

500

600

Particle size (nm)

(c)

Tr180t30Ts600

Number (%)

30

pH 11
20

pH 5

pH 9

pH 1

10

0
0

100 200 300 400 500 600 700 800

Particle size (nm)


Fig. 4. Particle size distributions of: (a) 1st set of experiments, (b) Tr130t30Ts600 at different pH values and (c) Tr180t30Ts600 at different pH values.

and synthesis time for the case of samples obtained below pH = 3.


As expected, the crystal sizes in experiments 7 (56 nm) and 8
(43 nm) were decreased after the sintering process when the initial
solution has the lowest pH value (1). From these comparisons, this
method led us to a more accurate value of L from all different
selected peaks (three most intense).
Therefore, it can be inferred that by adjusting the pH from 5 to
9, monoclinic type CePO4 nanostructures began to be formed,
whereas at pH 11, the hexagonal phase was transformed totally
to monazite in the synthesis independently of the reaction temperature (130 and 180 C) without needing further sintering
treatment.
The structure and morphology of selected CePO4 nanostructures, using both sets of experiments, were evaluated by the TEM
and SEM techniques as well as by selected area electron diffraction
patterns (SAED). In the rst set of experiments, by varying the reaction (Fig. 2a and b), the SAEDs corroborated the presence of two
structural phases at 400 C of sintering temperature, which grew
predominantly in nanorod shape and some particle-like agglomerate morphologies with a wide variety of sizes; rod nanostructures
consist of diameters from 5 to 12 nm and lengths up to 221 nm. By
increasing the sintering temperature to 600 C, the nanorods disappeared to form semi-spherical agglomerates with diameters below
20 nm (Fig. 2c and d). As expected, at 600 C, a strong relationship
between the sintering temperature and microstructure was
obtained.
Scanning electron microscope images of the progress of CePO4
nanostructures with the pH values (1, 5, 9 and 11) for the conditions used with experiments 7 and 8 are presented in Fig. 3ah,
respectively. Hexagonal nanorods of approximately 100 nm1 lm
in length and 10100 nm in diameter were obtained at pH 1
(Fig. 3a and e). At intermediate pH = 5, the morphology begins to
change and transforms into agglomerates of semi-spherical nanoparticles (Fig. 3b and f). Under these conditions, both morphologies

coexist: nanorods and agglomerated particles. By increasing the pH


up to 9, the predominant morphologies are spherical particles, but
thin and ne nanorods are also detected in some areas (Fig. 3c and
g). Interestingly, the nanorods become more evident by increasing
the reaction temperature (130 C). Finally, at pH = 11, the CePO4
nanopowders only consist of semi-spherical particles.
From these observations, it is clear that the morphological features of the CePO4 nanostructures can be controlled by the pH of
the reaction medium.
The signicance of a morphology-controlled synthesis and the
special properties of CePO4 nanopowders have been well acknowledged [18] due to the fact that different shapes often display different surface structures and always expose different active planes.
Their sizes, surface structures, and interparticle interactions can
develop some unique properties and improved performances for
future applications. Therefore, it is evident that the materials will
have a different performance as a function of the reaction medium.
Particle diameter moments (number-average diameter Dn;
weight-average diameter Dw and z-average diameter) were calculated using Eqs. (5)(7), and the polydispersity index (PDI) was
determined using Eq. (8):

Rni Di
Rni
Rni D4i
Dw
Rni D3i

Dn

Dz

5
6

Rni D6i
Rni D5i

Dw
Dn

PDI

where ni is the number of CePO4 nanoparticles with diameter Di.


The particle size distribution (PSD) graphs as well as average

Please cite this article in press as: D. Palma-Ramrez et al., J. Alloys Comp. (2015), http://dx.doi.org/10.1016/j.jallcom.2014.12.053

D. Palma-Ramrez et al. / Journal of Alloys and Compounds xxx (2015) xxxxxx

Transmittance (%)

r130

15

s400

955

564

578

1140-983

r130

30

r180

30

s400
538

800
-1

600

1400

1200

15

r130

30

T
s600

564

s600

r180

r180

1000

30

600

s600

537

1069
-1

s600

1091

800

15

616

955

995

Wavenumber (cm )

564

1069 1016

1200

600

578

955

1400

800

-1

537
616
1091

1000

Wavenumber (cm )

578
T

618

1140-983

1000

564

578

Transmittance (%)

Transmittance (%)

r130

s400

955

Wavenumber (cm )
T

618

s400

1200

(b)

15

1140-983

1140-983
1400

955

541
955

r180

Transmittance (%)

(a)

1400

1200

995
1016
1000

800

600

-1

Wavenumber (cm )

Fig. 5. FT-IR spectra of CePO4 nanopowders sintered at: (a) 400 C and (b) 600 C, under the conditions of the rst set of experiments.

particle size (Dz), polydispersity index (PDI), the highest mean


number (%) and sizes are shown in Table 4 and Fig. 4.
The results of the size distribution analyses show that the
majority of the samples feature a broad distribution. For example,
the average particle size of the rst set of experiments (at pH = 3) is
found between 541315 nm with PDI values above 1, indicating
the high polydispersity in the system. At pH = 1, the distribution
is slightly displaced to lower sizes; sample 7 showed two groups
at about 42 and 122 nm, conrming the mixed morphology seen
by TEM and SEM micrographs, while the nanopowders at pH = 5,
9 and 11 displayed the highest % of nanoparticles at 140
190 nm, 121531 nm, and 90125 nm, respectively. From these
observations and the associations with TEM and MEB, we can conclude that the powders consist of polydispersed agglomerates
which need to be separated into individual particles.
The effect of the microwave-assisted hydrothermal reaction
parameters on the CePO4 nanoparticles were also analyzed by

means of FT-IR, and the results of nanostructures treated at 400


and 600 C for the rst set of experiments in the 1500450 cm1
region are shown in Fig. 5a and b.
The IR spectra of the 14 and 58 materials synthesized under
the specied experimental conditions and sintered at 400 C are
presented in Fig. 5a, respectively. The spectra of the 14 conditions
display peaks at about 1140983 cm1 (antisymmetric stretching
vibration of the PO bond (m3)), 618, 578, 564, and 539 cm1
(asymmetric bending modes of the PO3
4 group (m4)) [19].
The FTIR spectra for samples 58 (Fig. 5b) show a similar tendency in the 950450 cm1 range. However, there are some differences between the analyzed sintering temperatures. One of the
most remarkable characteristics of the samples corresponding to
the 14 condition is observed at the 955 cm1 signal (symmetric
stretching vibrations). It is observed that the intensity of the peak
is more evident as the sintering temperature is increased from
400 to 600 C; this is due to the vibration of the PO bond in the

Please cite this article in press as: D. Palma-Ramrez et al., J. Alloys Comp. (2015), http://dx.doi.org/10.1016/j.jallcom.2014.12.053

D. Palma-Ramrez et al. / Journal of Alloys and Compounds xxx (2015) xxxxxx

OH-

OHLower ionic motion


Nucleation

CePO4

CePO4 spherical particles

PO 4
Ce

H+

CePO4 complex H+
Nucleation
Faster ionic motion

CePO4 rods like-particles


Fig. 6. Schematic representation of the formation of nanospherical particles and
nanorods.

(a)

100

130C, 15 min

8
5

80

180C 30 min

7,4

180C, 15 min

60

1,3
2
130C 30 min
Diffuse reflectance (%)

Diffuse reflectance (%)

40
20
0

20

68
5

15

7
10

180C 30 m in
180C, 15 m in
4
130C 30 m in
2 1
3

H5 P3 O10 aq H2 Oaq ! H4 P2 O7 aq H3 PO4 aq

0
200

H4 P2 O7 H2 O ! 2H3 PO4

220

240

260

280

300

(nm)

4HNO3 aq nH2 Oaq

7(pH 1)

8(pH 1)
8(pH 9)

40

7(pH 9)

20
7(pH 11)

0
200

Diffuse reflectance (%)

Diffuse reflectance (%)

7(pH 5)
8(pH 5)

60

8(pH 11)

300

400
500
(nm )

25
20
15
10
5

7 (p H 1 )

7 (p H 9 )

8 (p H 9 )
8 (p H 5 )
7 (p H 9 )
8 (p H 1 )
8 (p H 1 1 )

7 (p H 1 1 )
200
220

240
260
(n m )

600

11

On the other hand, the main reaction might take place according to
Eqs. (12)(14). The precipitates of the Ce3+PO3
4 complex were more
evident when the pH of the initial solution was adjusted in alkaline
medium. This precipitate leads to a lower concentration of Ce3+ and

HnPO3
4 , decreasing the chemical potential, and the OH ions from
NH4OH are absorbed on the crystal face, altering the growth of
the nanorods to form spherical particles belonging to the monoclinic system as it is also shown in Fig. 6.

(b)

80

9
10

CeNO3 3 aq  6H2 O HNO3 aq H3 PO4 aq ! CePO4 s #

200 250 300 350 400 450 500 550 600 650 700
(nm)
100

The formation of nanomaterials, nucleation rates and growth


process are inuenced by many reaction conditions such as pH,
pressure, temperature, precursor concentration, solvent [21]. In
such a case, the controlled addition of precursors or surfactants
or templates is often used as a feasible means to control the nucleation rates and growth process, thus reaching the aim of controlling the crystal size. However, the synthesis method is
3
particularly important. Ions such as NO
3 ; PO4 come mainly from
the cerium salt precursor and the precipitant [18]. These anions
have a selective interaction with specic facets [22], which the
anions present during the synthesis of CePO4 nanomaterials via
the hydrothermal/solvothermal method, which leads to the growth
of nanocrystals with varying morphologies. In such a case, the formation mechanism of nanorods or spherical particles depends on
the acid or alkaline medium as follows:
In the case of acid medium, the hydrothermal reaction system
consists of Ce(NO)3H2O, H2O, HNO3 and H3PO4 coming from the
dissolution of H5P3O4. The mechanism most probably involves
the release of PO3
4 ions present in the H3PO4 solution, which are
combined with Ce3+ ions to form an amorphous precipitate in an
aqueous solution. During the experimental work, it was observed
that as the pH value was decreased, the Ce3+PO3
complex was
4
more dissolved. According to Y. Zhang, this dissolution behavior
results in a faster ionic motion in the solution, leading to a higher
concentration of Ce3+ and HnPO3
4 , where the chemical potential is
increased [23] and promotes the preferential adsorption of H+ onto
the crystal facets, which seems to raise the electrostatic potential
on the crystal surfaces [6]. Thus, the preferential adsorption of H+
leads to the hexagonal nanorod morphology as shown in Fig. 6.
This process can be described with Eqs. (9)(11).

280

300

700

Fig. 7. Diffuse reectance spectra obtained for all the as-prepared powders, before
and after the sintering process.

monoclinic structure (m1), indicating the rearrangement of the


hexagonal structure to form the monoclinic one. These facts
corroborate and conrm the observations made in the XRD patterns for the samples sintered at 400 C, where the transformation
of rhabdophane was observed.
It is worth mentioning that in addition to this change, there is a
remarkable difference in the broad band region between 1140 and
983 cm1, where it is observed how the peaks start to be more
dened and evident at about 1091, 1069, 1016 and 995 cm1
(asymmetric stretching of PO bonds). This behavior indicates a
complete rearrangement and it has been observed by calcination
studies at different temperatures by Pusztai et al. [20].

H5 P3 O10 aq H2 Oaq ! H4 P2 O7 aq H3 PO4 aq

12

H4 P2 O7 H2 O ! 2H3 PO4

13

CeNO3 3 aq  6H2 Oaq NH4 OHaq H3 PO4 aq


! CePO4 s # 2NH4 NO3 aq nH2

14

3.2. Optical properties of CePO4 nanostructures


The optical absorption properties of the CePO4 materials were
characterized by diffuse reectance (DR) measurements. Fig. 7a
and b shows the diffuse reectance spectra of CePO4 nanoparticles
obtained under the conditions of the aforementioned rst and second sets of experiments. The measurements were carried out
before and after the sintering process in the ultraviolet light range,
which can be divided into three sub-intervals: the UV-C (100
280 nm), UV-B (280315 nm), UV-A (315400 nm) and visible
light (400700 nm).
The frequent use of diffuse reectance spectroscopy for the
determination of the electronic transitions in solid materials is
highly justied due to the optical excitation of the electrons from
the valence band to the conduction band, which is evidenced by

Please cite this article in press as: D. Palma-Ramrez et al., J. Alloys Comp. (2015), http://dx.doi.org/10.1016/j.jallcom.2014.12.053

D. Palma-Ramrez et al. / Journal of Alloys and Compounds xxx (2015) xxxxxx

Tr130t15Ts600

Tr130 t30Ts600

pH 3
253

300

(F(R)hv) 2

225

(F(R)hv) 2

(F(R)hv) 2

(F(R)hv) 2

pH 1

200 250 300 350 400 450 500

257 273

200

(nm)

240

215
220

240

3.22 eV
1.0

1.5

2.0

2.5

3.5

4.0

1.0

1.5

2.0

273

240

220

240

260

280

(nm)

(F(R)hv) 2

(F(R)hv) 2

pH 5

(F(R)hv) 2

(F(R)hv) 2

Tr130t30Ts600

pH 3

200

1.5

2.0

300

224

3.04 eV

3.0

3.5

4.0

1.0

1.5

2.0

257 273
240
215

280

200

300

220

2.5

3.5

4.0

240

260 280

(nm)

300

3.25 eV

3.23 eV

2.0

3.0

(F(R)hv) 2

(F(R)hv) 2

(F(R)hv) 2

pH 11

(F(R)hv) 2

Tr130t30Ts600

220 240 260

2.5

hv (eV)

Tr180t30Ts600

1.5

4.0

200 220 240 260 280 300

pH 9

1.0

3.5

(nm)

2.5

(nm)

3.0

266 281

hv (eV)

200

2.5

250

3.0 eV
1.0

300

hv (eV)

Tr130t15Ts400

215

280

3.06 eV
3.0

hv (eV)

257

260

(nm)

3.0

3.5

4.0

1.0

1.5

2.0

2.5

3.0

3.5

4.0

hv (eV)

hv (eV)

Fig. 8. KubelkaMunk modied spectra and its equivalence in hm vs k (inset) of selected CePO4 samples.

an increase in the absorbance at a given wavelength (band gap


energy) [24].
In general, from Fig. 7ab two main features were observed in
the CePO4 powders during the diffuse reectance study: rst, all
the spectra revealed large absorption in the UV-B (280315 nm)
and UV-A (315400 nm) regions with high transparence under
the visible light range (400700 nm), attributable to the weak
absorption in the as-prepared nanostructures. This region can be
particularly important for practical applications in ultravioletlight-based-white-light-emitting dioxides (UV-WLED), components for precision optical devices, windows, etc. [25]. A second
evident feature in the samples is the increase in the absorption
with the hexagonal structure (at 400 C), which is a desired condition for polymer applications since UV-A and UV-B lights are
responsible for color loss, mechanical property alterations and
structural modications [26].
The quantication of CePO4 nanopowders indicated that after
the sintering process at 400 C, the samples display higher reectance percent in the 400700 nm region. On the other hand, CePO4
nanostructures sintered at 600 C show a tendency similar to that
of the non-sintered powders. In comparison, the pH modied samples (Fig. 7b) show high absorption characteristics in the UV interval, while from all these samples, samples 7 and 8 reect the best

stability in the visible region at pH = 1. Additionally, it was


observed that CePO4 structures obtained at pH = 11 displayed the
best absorption properties. As expected, the reectance can be
greatly modied by the structural characteristics of the CePO4
nanoparticles i.e., shape, size and roughness [27].
An approximation of the optical absorption coefcient was
evaluated by the KulbekaMunk function, using the reectance
data to calculate the absorption coefcient (F(R1)),

F R1

1  R1 2 Kk

sk
2R1

15

where (F(R1)) is the MK function or re-emission function, R1 is the


diffused reectance at a given wavelength of an innitely thick
sample, K(k) is the absorption coefcient, and s(k) is the scattering
coefcient. For direct band gap transitions, the optical band
gap
value
(Eg)
is
calculated
using
the
relation
1=2
FR1hm Ahm  Eg , where A is constant, hm is the photon
energy and Eg is the optical band gap. In all the cases, Eg is inferred
by extrapolating linearly the long-wave-length edge of the peak in
2
absorbance to this zero line from the plot between FR1hm and
(hm) [28,29].
The band gap was calculated from the linear part of the slope.
Selected results for the overall experiments are presented in

Please cite this article in press as: D. Palma-Ramrez et al., J. Alloys Comp. (2015), http://dx.doi.org/10.1016/j.jallcom.2014.12.053

10

D. Palma-Ramrez et al. / Journal of Alloys and Compounds xxx (2015) xxxxxx

Fig. 8, whereas a summary of the entire samples is presented in


Table 5. In the inset gures, the spectrum of condition 7 obtained
at pH = 1 shows remarkable similarities with respect to the absorption spectra of bulk CePO4 reported by Yue-Ping Fang [30]. The
FR1hm2 vs. k (nm) curves can also be seen, where it was found
that within pH 19 appears a region (200300 nm) with four peaks
that corresponds to the transitions of the fd orbitals from the
ground state 2F5/2(4f1) of Ce3+ to the ve crystal eld split levels
of the Ce3+ 2D(5d1) excited states [30]; the small displacement at
pH = 5 was correlated with differences in the particle size. It is
important to note that these transitions cannot be resolved in
the spectra when the solution pH is adjusted to 11.
It can be mentioned that the results evidence that the formation
of CePO4 powders in the hexagonal or monoclinic phase are highly
dependent on the combination of different factors, however, the
constant microwave energy, sintering temperature and pH seem
to be the most important parameters. A structural stability of the
hexagonal or monazite phases can be reached by using a microwave-assisted hydrothermal method at relatively low process temperatures by adjusting the solution pH. Finally, the study
emphasizes the clear relation among the process conditions, structure and sample morphologies, which in turn reected changes in
the optical properties.
4. Conclusions
An enhanced hydrothermal synthesis by using microwave
energy for obtaining CePO4 nanostructures was reported. The
method demonstrated to be a fast, low energy and capable one
in the synthesis of CePO4 nanostructures, and an option to replace
the commonly used method, which needs long periods of time to
reach a nanoscale. The present work also showed the strong relation between pH and the nal structural properties of cerium phosphate nanoparticles. By evaluating the structural and
morphological properties, it was observed that by means of high
microwave energy (200 W), a combination of hexagonal nanorods
and monoclinic nanoparticles can be achieved at 400 C, whereas
pure monazite was obtained by sintering the samples at 600 C.
By adjusting the reaction solution pH, nanorods are favored at
pH = 1, whereas a very alkaline medium (pH = 11) caused the formation of CePO4 semispherical agglomerates. The results highlighted the dependence of the structure on the pH, where
monazite was formed under mild conditions. Finally, the low diffuse reectance values in the UV region indicate that the nanopowders can be used as llers to increase the UV resistance in
polymeric materials.
Acknowledgments
D. Palma-Ramrez is grateful for her postgraduate scholarship
by SIP-IPN and COFAA-IPN. The authors are also grateful for the
nancial support provided by CONACYT Mxico through the
CB2009-132660 and CB2009-133618 projects and by IPN through
the SIP 2014-0164 and 2014-0992 projects and SNI-CONACYT
Mxico.
References
[1] R. Yang, J. Qin, M. Li, Y. Liu, F. Li, Redox hydrothermal synthesis of cerium
phosphate microspheres with different architectures, CrystEngComm 13
(2011) 72847292.
[2] J. Bao, R. Yu, J. Zhang, X. Yang, D. Wang, J. Deng, J. Chen, X. Xing, Lowtemperature hydrothermal synthesis and structure control of nano-sized
CePO4, CrystEngComm 11 (2009) 16301634.

[3] L. Ma, W.-X. Chen, Y.-F. Zheng, Z.-D. Xu, Hydrothermal growth and morphology
evolution of CePO4 aggregates by a complexing method, Mater. Res. Bull. 43
(2008) 28402849.
[4] P. Xu, R. Yu, L. Zong, J. Wang, D. Wang, J. Deng, J. Chen, X. Xing, Phase evolution
and photoluminescence enhancement of CePO4 nanowires from a low
phosphate concentration system, J. Nanoparticle Res. C7-1622 15 (2013) 18.
[5] I. Bilecka, M. Niederberger, Microwave chemistry for inorganic nanomaterials
synthesis, Nanoscale 2 (2010) 13581374.
[6] N. Ekthammathat, T. Thongtem, A. Phuruangrat, S. Thongtem, Microwaveassisted synthesis of CePO4 nanorod phosphor with violet emission, Rare Met.
30 (2011) 572576.
[7] C.R. Patra, G. Alexandra, S. Patra, D.S. Jacob, A. Gedanken, A. Landau, Y. Gofer,
Microwave approach for the synthesis of rhabdophane-type lanthanide
orthophosphate (Ln = La, Ce, Nd, Sm, Eu, Gd and Tb) nanorods under
solvothermal conditions, New J. Chem. 29 (2005) 733739.
[8] L. Karpowich, S. Wilcke, R. Yu, G. Harley, J.A. Reimer, L.C. De Jonghe, Synthesis
and characterization of mixed-morphology CePO4 nanoparticles, J. Solid State
Chem. 180 (2007) 840846.
[9] L. Macalik, P.E. Tomaszewski, A. Matraszek, I. Szczygiel, P. Solarz, P. Godlewska,
M. Sobczyk, J. Hanuza, Optical and structural characterisation of pure and Pr3+
doped LaPO4 and CePO4 nanocrystals, J. Alloys Comp. 509 (2011) 74587465.
[10] N. Clavier, R. Podor, N. Dacheux, Crystal chemistry of the monazite structure, J.
Eur. Ceram. Soc. 31 (2011) 941976.
[11] S. Lucas, E. Champion, D. Bregiroux, D. Bernache-Assollant, F. Audubert, Rare
earth phosphate powders RePO4nH2O (Re = La, Ce or Y) Part I. Synthesis and
characterization, J. Solid State Chem. 177 (2004) 13021311.
[12] N. Ekthammathat, T. Thongtem, A. Phuruangrat, S. Thongtem, Microwaveassisted synthesis and characterisation of uniform LaPO4 nanorods, J. Exp.
Nanosci. 7 (2012) 616623.
[13] H. Hejase, S.S. Hayek, S. Qadri, Y. Haik, MnZnFe nanoparticles for selfcontrolled magnetic hyperthermia, J. Magn. Magn. Mater. 324 (2012) 3620
3628.
[14] J.I. Langford, A.J.C. Wilson, Scherrer after sixty years: a survey and some new
results in the determination of crystallite size, J. Appl. Crystallogr. 11 (1978)
102113.
[15] A. Monshi, M.R. Foroughi, M.R. Monshi, Modied Scherrer equation to estimate
more accurately nano-crystallite size using XRD, World J. Nano Sci. Eng. 2
(2012) 154160.
[16] P.S. Nair, T. Radhakrishnan, N. Revaprasadu, G.A. Kolawole, P. OBrien, A singlesource route to CdS nanorods, Chem. Commun. (2002) 564565
[17] A. Gaber, M.A. Abdel- Rahim, A.Y. Abdel-Latief, M.N. Abdel-Salam, Inuence of
calcination temperature on the structure and porosity of nanocrystalline SnO2
synthesized by a conventional precipitation method, Int. J. Electrochem. Sci. 9
(2014) 8195.
[18] D. Zhang, X. Du, L. Shi, R. Gao, Shape-controlled synthesis and catalytic
application of ceria nanomaterials, Dalton Trans. 41 (2012) 1445514475.
[19] T. Masui, H. Tategaki, S. Furukawa, N. Imanaka, Synthesis and characterization
of new environmentallyfriendly pigments based on cerium phosphate, J.
Ceram. Soc. Jpn. 112 (2004) 646649.
[20] P. Pusztai, T. Simon, . Kukovecz, Z. Knya, Structural stability test of
hexagonal CePO4 nanowires synthesized at ambient temperature, J. Mol.
Struct. 1044 (2013) 9498.
[21] Q. Yuan, H.-H. Duan, L.-L. Li, L.-D. Sun, Y.-W. Zhang, C.-H. Yan, Controlled
synthesis and assembly of ceria-based nanomaterials, J. Colloid Interface Sci.
335 (2009) 151167.
[22] W. Wang, J.Y. Howe, Y. Li, X. Qiu, D.C. Joy, M.P. Paranthaman, M.J. Doktycz, B.
Gu, A surfactant and template-free route for synthesizing ceria nanocrystals
with tunable morphologies, J. Mater. Chem. 20 (2010) 77767781.
[23] Y. Zhang, H. Guan, The growth of lanthanum phosphate (rhabdophane)
nanobers via the hydrothermal method, Mater. Res. Bull. 40 (2005) 1536
1543.
[24] R. Lpez, R. Gmez, Band-gap energy estimation from diffuse reectance
measurements on solgel and commercial TiO2: a comparative study, J. Sol
Gel Sci. Technol. 61 (2011) 17.
[25] Y.-Q. Li, S.-Y. Fu, Y.-W. Mai, Preparation and characterization of transparent
ZnO/epoxy nanocomposites with high-UV shielding efciency, Polymer 47
(2006) 21272132.
[26] A.L. Andrady, H. Hamid, A. Torikai, Effects of solar UV and climate change on
materials, Photochem. Photobiol. Sci. 10 (2011) 292300.
[27] S. Dhara, P.K. Giri, Enhanced UV photosensitivity from rapid thermal annealed
vertically aligned ZnO nanowires, Nanoscale Res. Lett. C7-504 6 (2011) 18.
[28] A. Chithambararaj, A.C. Bose, Hydrothermal synthesis of hexagonal and
orthorhombic MoO3 nanoparticles, J. Alloys Comp. 509 (2011) 81058110.
[29] A.B. Murphy, Band-gap determination from diffuse reectance measurements
of semiconductor lms, and application to photoelectrochemical watersplitting, Sol. Energy Mater. Sol. Cells 91 (2007) 13261337.
[30] Y.-P. Fang, A.-W. Xu, R.-Q. Song, H.-X. Zhang, L.-P. You, J.C. Yu, H.-Q. Liu,
Systematic synthesis and characterization of single-crystal lanthanide
orthophosphate nanowires, J. Am. Chem. Soc. 125 (2003) 1602516034.

Please cite this article in press as: D. Palma-Ramrez et al., J. Alloys Comp. (2015), http://dx.doi.org/10.1016/j.jallcom.2014.12.053

Вам также может понравиться