Você está na página 1de 12

Journal of Alloys and Compounds 685 (2016) 848e859

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Morphology evolution of Gd0.99Eu0.01BO3 phosphors synthesized by


hydrothermal method with Gd0.99Eu0.01(OH)3 nanorods as sacricial
precursors
Yingli Zhu a, b, Yujun Liang a, b, *, Shiqi Liu a, b, Kai Li a, b, Xingya Wu a, b, Rui Xu a, b
a
b

Engineering Research Center of Nano-Geomaterials of Ministry of Education, China University of Geosciences, Wuhan 430074, Peoples Republic of China
Faculty of Materials Science and Chemistry, China University of Geosciences, Wuhan 430074, Peoples Republic of China

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 12 May 2016
Received in revised form
12 June 2016
Accepted 21 June 2016
Available online 23 June 2016

Monodisperse mono-concave microsphere/microower/nanosheet GdBO3:Eu3 samples have been


successfully synthesized via hydrothermal method using Gd0.99Eu0.01(OH)3 nanorods as sacricial precursors. X-Ray diffraction (XRD), Fourier transform infrared spectroscopy (FT-IR), eld scanning electron
microscopy (FE-SEM) and uorescence spectrometry were used to characterize the samples. The shape of
these as-prepared GdBO3:Eu3 architectures can be tuned effectively by controlling the concentrations of
NaOH (CNaOH) of the precursor solution and the amounts of PEG (20000) surfactant. As a typical
morphology, the possible formation mechanism for GdBO3:Eu3 monodisperse mono-concave microsphere was presented in detail. Under ultraviolet excitation, the as-obtained GdBO3:Eu3 samples
showed strong characteristic orange and red emission of Eu3. Compared with samples get by solid state
reaction, the samples obtained by hydrothermal method possess higher red/orange (R/O) ratio attributed
to 5D07F1 and 5D07F2 transitions. The inuence of diversity morphology on the R/O ratio and emission
intensity has been also investigated. This work sheds some light on the design of a well-dened complex
structure, and the luminescent properties have potential applications in uorescent lamps and eld
emission displays.
2016 Elsevier B.V. All rights reserved.

Keywords:
Phosphors
Liquid-solid reactions
Optical properties
Scanning electron microscopy
SEM
Luminescence

1. Introduction
Nowadays, development of modern technology requires the
creation of advanced materials with suitable physicochemical
properties. One of the most intensively developing technology
elds is related to the synthesis and utilization of nano/microstructures with well-dened and controllable shapes, which are
expected to be useful in a wide range of applications, such as
catalysis, nano-electronics, optics and biosensors [1e5]. Their
chemical and physical properties are closely associated with not
only their geometrical factors such as size, dimensionality,
morphology, but also the assembly driving force, orientation and
assembly pattern [6]. Recently, chemical self-assembly has become
an important method in many elds. And what is more, it is

* Corresponding author. Engineering Research Center of Nano-Geomaterials of


Ministry of Education, China University of Geosciences, Wuhan 430074, Peoples
Republic of China.
E-mail address: yujunliang@sohu.com (Y. Liang).
http://dx.doi.org/10.1016/j.jallcom.2016.06.215
0925-8388/ 2016 Elsevier B.V. All rights reserved.

accepted as one of the top 25 big questions facing science over the
next quarter-century and the only practical approach for building a
wide variety of nanostructures. The self-assembly of building
blocks during their growth or mature stage can provide another
approach to develop the sophisticated nano/microarchitectures
[7e11]. To date, many complex nano/microstructures have been
synthesized based on the self-assembly process, such as ZnS
nanorods [12], CuO ellipsoids [13], CdTe microsheets [14], MoS2
microbundles [15]. Researchers have paid great attention to
investigate the law of self-assembly in self-assembling systems.
However, the discipline of the self-assembly in complex structure
has never been made clear. So far, scientists are still exploring the
rule of the self-assembly.
Its well known that the polymorphism of rare-earth orthoborates LnBO3 (Ln lanthanides and yttrium) has produced
numerous studies concerning their crystallographic structures and
chemical properties [16e19]. Firstly, the high structural exibility
of these solids is caused by the junction of planar and nonplanar
BO3 groups and BO4 tetrahedra, which can occur as isolated or
compact fundamental building units. In 1961 Levin et al.

Y. Zhu et al. / Journal of Alloys and Compounds 685 (2016) 848e859

characterized three elementary crystal structures of LnBO3 as three


crystalline forms of CaCO3, which are associated with borates
preparation temperature and the ionic radius of the rare earth. The
LnBO3 of the light rare earths (La to Nd) exhibit the aragonite-type
structure and those of the heavy rare earths (Sm to Yb and Y)
possess the vaterite-type structure. Other class, LuBO3 and ScBO3
are known to form in the calcite-type structure [20,21]. In addition,
LnBO3 materials are applied for the color plasma display panels and
Hg-free uorescent lamps in view of their high stability, low synthesis temperature, high ultraviolet and exceptionally optical
damage threshold. It is found that in rare-earth-doped LnBO3, the
luminescence properties are related to the size, shape, and crystal
types, which can be governed by synthetic routes and parameters
[22,23]. So the present tendency of developing such materials is not
focusing on especially novel crystalline phases but mostly on a
better control of microstructures of existing materials. Among
LnBO3 materials, GdBO3, which processes the hexagonal vateritetype structure and shows a good vacuum ultraviolet (VUV) absorption, has been a promising host material for doping Eu3 VUV
phosphors [24,25]. However, GdBO3:Eu3 phosphors derived from
conventional solid-state reaction have irregular morphology, which
will limit its dispersity and subsequent coating ability on the
display panels. Therefore, it is greatly desired to synthesize
GdBO3:Eu3 phosphor with homogeneous and adjustable
morphology by a controllable approach. The hydrothermal
approach is a good choice due to its convenience, exemption from
pollution, and the possibility of achieving satisfying crystallinity at
a relatively low temperature, without further calcination [26,27].
Up to now, various morphologies of Eu3 or Tb3 doped GdBO3
have been synthesized via hydrothermal method, such as spindleassembled nanocrystals [28], uniform microplates [16], microowers architectures [29], rodlike nanoparticles [24], 2D nanosheets architectures [30], 1D nanobres [31] and so on. However, to
our knowledge, few reports have been focused on the control of
diverse morphologies of GdBO3 by using Gd0.99Eu0.01(OH)3 as
precursors.
Recently, many sacricial precursors have been used to fabricate
different morphologies of lanthanide compounds, such as Ln(OH)
CO3 (Ln Y, Gd, Yb and Lu) [32e34], Lu4O(OH)9NO3 [35], etc. It is
also noteworthy that Gd(OH)3 nanorods or nanowires as sacricial
precursors have been successfully used to fabricate different kinds
of lanthanide compounds [29,30,36]. Here in our paper, nearly
monodisperse self-assembly GdBO3:Eu3 phosphors with various
morphologies by using Gd0.99Eu0.01(OH)3 nanorods as sacricial
precursors have been successfully prepared through a mild hydrothermal method. First, we prepared Gd0.99Eu0.01(OH)3 nanorods
through a facile solution-based hydrothermal process without using any catalyst or template. Then, Gd0.99Eu0.01(OH)3 nanorods as
both the physical and chemical precursors were used to prepare
GdBO3:Eu3 architecture. The shape and dimensionality of the
obtained GdBO3:Eu3 architectures can be controlled conveniently
by reactive basicity and PEG (20000) amounts. Furthermore, the
structure, formation mechanism and photoluminescence properties of GdBO3:Eu3 products have also been investigated in detail.
Results show that this research is signicant for preparing and
exploiting potential applications of LnBO3 manifold structures in
many elds: nanoapparatus, optics, catalysis, intercalation chemistry, solid batteries, and so forth.
2. Experimental
2.1. Materials
Gd2O3, Eu2O3 (99.99%) were purchased from Science and
Technology Parent Company of Changchun Institute of Applied

849

Chemistry, and other chemicals were purchased from Beijing Fine


Chemical Company. All chemicals were analytical-grade reagents
and used as purchased without further purication.
2.2. Hydrothermal synthesis of Gd0.99Eu0.01(OH)3 nanorods
The Gd0.99Eu0.01(OH)3 nanorods precursors were prepared via a
typical hydrothermal synthetic process. 5 mmol Gd(NO3)3 and
Eu(NO3)3 solution (20 mL, 0.25 M) were measured and dissolved in
deionized water under continuous stirring. Subsequently, 10 mL
NaOH aqueous solution was introduced into the solution with
concentration of NaOH ranging from 1.0 to 5.0 M. After further
stirring, the resultant mixture was subsequently transferred into a
50 mL Teon-lined autoclave. The hydrothermal reaction was
conducted at 200  C for 24 h. After the autoclave was cooled to
room temperature, the obtained white products were washed with
distilled water several times.
2.3. Hydrothermal syntheses of GdBO3:Eu3
In a typical procedure for the preparation of GdBO3:Eu3, the asobtained 1 mmol Gd0.99Eu0.01(OH)3 nanorods were rstly redispersed into distilled water by ultrasonic treatment.
10 mL H3BO3 (0.5 M) aqueous solution was dripped into the
dispersed solution followed by further stirring. The mixed solution
was then transferred to a 50 mL Teon-line autoclave and sealed.
The autoclave was heated to 200  C and maintained at this temperature for 24 h. After naturally cooling down to room temperature, the obtained white products was washed with distilled water
and ethanol and then dried at 80  C. As a contrast, the GdBO3:Eu3
samples were also synthesized by a one-step hydrothermal method
without utilizing Gd0.99Eu0.01(OH)3 nanorods as precursors. As raw
materials, 1 mmol of Gd(NO3)3 and Eu(NO3)3 solution (4 mL,
0.25 M) were measured and dissolved in deionized water under
continuous stirring. Then, the stoichiometric amounts of H3BO3
were dissolved in the solution. Subsequently, NaOH aqueous solution was introduced into the solution until pH 12.0. After additional agitation for 30 min, the mixed solution was transferred to an
autoclave, and hydrothermal treatment was performed at 200  C
for 24 h. Finally, the samples were obtained according to above
processing method.
2.4. Solid state reaction synthesis of GdBO3:Eu3
For comparison, the GdBO3:Eu3 samples were also synthesized
using a solid state reaction. As raw materials, the stoichiometric
amounts of Gd2O3, Eu2O3 and H3BO3 were weighed and mixed well
in an agate mortar. The mixture was then transferred to an alumina
crucible and calcined at 1300  C for 4 h in an oxygen atmosphere.
Finally, the as-obtained samples were slowly cooled to room temperature for subsequent characterization.
2.5. Characterizations
X-ray powder diffraction (XRD) patterns of the materials were
recorded using a Bruker D8 Focus diffractometer with Ni-ltered
Cu-Ka (l 1.540598 ) radiation at 40 kV tube voltage and
40 mA tube current. The morphologies of the obtained samples
were characterized by a Japan SU8010 eld emission scanning
electron microscopy (FE-SEM) at 15 kV. Excitation and emission
spectra were recorded on a Horiba Jobin Yvon FLUOROMAX-4P
uorescence spectrometer equipped with a 150 W xenon lamp.
The spectral step length of both the excitation and emission spectra
was set up to be 1.0 nm with the width of the monochromator slits
adjusted to 0.50 nm. The other measurement conditions were kept

850

Y. Zhu et al. / Journal of Alloys and Compounds 685 (2016) 848e859

consistent from sample to sample during the measurements. All the


measurements were carried out at room temperature.
3. Results and discussion
3.1. Phase identication and morphology of the Gd0.99Eu0.01(OH)3
precursors
Fig. S1 shows the X-ray diffraction patterns of the precursors
Gd0.99Eu0.01(OH)3 obtained by treating Gd(NO3)3 with different
CNaOH, i.e., 1.0, 2.0, 3.0, 4.0 and 5.0 M at 200  C for 24 h. It was
observed that under different CNaOH, all diffraction peaks of the
precursors could be well indexed to the hexagonal phase (space
group: P63/m (No. 176)) of Gd(OH)3 (JCPDS No. 83-2037) with lattice constants a b 6.329 , c 3.631 , V 125.96 3. The CNaOH
of the colloidal solution had certain inuence on the size and
morphology of the hydrothermal products Gd0.99Eu0.01(OH)3.
Fig. S2 reveals the panoramic SEM images of the Gd0.99Eu0.01(OH)3
precursors obtained under different CNaOH (CNaOH 1.0 M, 2.0 M,
3.0 M, 4.0 M, 5.0 M), from which we could see that the as-prepared
samples were almost entirely composed of nanorods at different
sizes. The diameter and length of the nanorods were range from
less than 50 nm to 300 nm and from less than 200 nm to more than
1 mm. When the CNaOH was relatively low, the amount of the pony
and large-size nanorods was almost the same. As the CNaOH
increased, the quantity of the pony-size nanorods gradually
decreased followed by the large-size nanorods increased. The size
of the nanorods became the most uniform when the CNaOH was up
to 5.0 M. Continue to increase the CNaOH, a few pony-size nanorods
appeared.

completely into the GdBO3 phase under the current synthetic


conditions. However, the intensities of the diffraction peaks varied
signicantly, which revealed different crystallinity and preferential
growth [37,38]. The samples prepared by one step method and solid
state reaction showed better crystallization than those prepared by
two steps method. In addition, it was worthy to note that in two
steps method, the relative intensity ratio of the diffraction peaks
varied obviously with the increased of the CNaOH value of the colloid
solution. It was found that the facets of (002) and (004) became
stronger, especially (002) facets, while the facets of (100), (102) and
(112) decreased greatly. In particular, the diffraction intensity ratio
of the (002)/(100) planes in the as-obtained GdBO3:Eu3 crystal
when CNaOH 5 M, increased 14.5 fold compared with the standard
value of 0.46. These results implied that the crystalline anisotropy
appeared due to the preferential orientation of a specic crystal
plane under hydrothermal environment. The effect of preferential
orientation on the morphology of the GdBO3:Eu3 crystal would be
spelt out exactly in Section 3.3.
In order to further understand the microstructure of the asprepared GdBO3:Eu3 crystals, detailed Rietveld renements
were performed on the samples of CNaOH 1 M and 5 M. The ICSD27934, space group P63/mmc structure was introduced as an initial

3.2. Phase identication, XRD renement, crystal structure and


morphology of GdBO3:Eu3 crystal
Fig. 1 displays the typical XRD patterns for the as-prepared
GdBO3:Eu3 crystal synthesized by Gd0.99Eu0.01(OH)3 precursors
at different CNaOH and obtained by one step method and solid state
reaction. All diffraction peaks of the products could be well indexed
to the pure hexagonal vaterite phase of GdBO3 (ICSD-27934, space
group P63/mmc), and no trace of characteristic peaks from other
impurity phases was observed, suggesting that this hydrothermal
method was a feasible route to prepare pure phase of GdBO3 crystal.
It also indicated that the Gd(OH)3 phase could transform

Fig. 1. XRD patterns of the as-prepared GdBO3:Eu3 phosphors synthesized by


Gd0.99Eu0.01(OH)3 precursors at the different CNaOH (CNaOH 1.0 M, 2.0 M, 3.0 M, 4.0 M,
5.0 M) and obtained by one step method and solid state reaction.

Fig. 2. Experimental (crosses) and calculated (red solid line) XRD patterns and their
difference (blue solid line) for the Rietveld t of GdBO3:Eu3 samples when
CNaOH 1 M (a) and 5 M (b). (For interpretation of the references to color in this gure
legend, the reader is referred to the web version of this article.)

Y. Zhu et al. / Journal of Alloys and Compounds 685 (2016) 848e859

model in the Rietveld analysis. Fig. 2 exhibits the Rietveld t of the


XRD pattern for GdBO3 crystals when CNaOH 1 M and 5 M. The
nal rened results of unit cell parameters and reliability factors for
the as-prepared samples are listed in Table 1, which are from
rened convergence and satised well with the reection condition. It was found that the cell volume of the two samples increased
with the introduction of Eu3 than the undoped 113.67 3 which
could be ascribed to the different ionic radii of the dopant ions
compared with those of the GdBO3 host. And the cell volume of the
sample when CNaOH 1 M was slightly larger than the CNaOH 5 M.
However, the c/a value of the two samples was almost equal,
indicating that when CNaOH was relative high, the preferential
orientation of the GdBO3 crystals had a little impact on the cell
parameters. All atom positions, fraction factors, occupation probability, thermal vibration parameters of GdBO3:Eu3 crystal when
CNaOH 1 M are showed in Table 2.
Generally, calcite, aragonite, or vaterite are three isostructural
forms of LnBO3. Depending on the size of lanthanide cations, their
orthoborates usually exhibit the aragonite-type (LaeNd), vateritetype (SmeYb), or calcite-type (Lu) structure [20,21]. Both the
calcite-type and the aragonite-type borates involve triangular
planar BO3
3 groups, but the vaterite-type borates involve ringy
9
B3O9
9 groups [39]. In order to conrm the presence of plane B3O9
rings in the GdBO3 crystals, the FT-IR spectra of the as-synthesized
GdBO3 crystals are shown in Fig. 3. The presence of several strong
absorption bands could be clearly detected. The intense absorption
bands in the region 750e1250 cm1 (centered at around 831, 903,
and 1169 cm1) could be ascribed to the characteristic vibration
modes of B3O9
9 groups in the vaterite-type orthoborates [18,19].
The vibration modes of B3O9
involved a planar ring with D3
9
symmetry. Two strong peaks at around 831 and 903 cm1 were
ascribed to the ring stretching vibration modes, whereas the peak
at around 1169 cm1 was ascribed to a terminal BeO stretching
vibration mode. While the presence of triangular BO3
3 groups in
the calcite-type and aragonite-type orthoborates were characterized by strong absorption band at 1290 cm1. The bands in the
range of 3200e3750 cm1 for the samples should be assigned to
the absorbed H2O in samples. On the basis of the XRD patterns and
FT-IR spectrum analysis, it was certain that the vaterite-type
GdBO3:Eu3 crystals were successfully obtained via the hydrothermal route by using Gd0.99Eu0.01(OH)3 nanorods as precursors.
Fig. 4(a) presents schematically the crystal structure of hexagonal GdBO3 crystals. The B3O9
9 groups were located on a 3-fold
screw axis, and in this way, the oxygen and boron atoms were no
longer partially occupied. A simple approach to understand this
structure was to stack B3O9
9 sheets, and the gadolinium atoms
were padded in the interlayer. Fig. 4(b) shows a sheet composed of
B3O9
9 and gadolinium atoms. The sheets were stacked in a rhombohedral fashion, forming a three-dimensional structure. The

Table 1
Final rened structure parameters of GdBO3:Eu3 samples using precursors
Gd0.99Eu0.01(OH)3 under CNaOH value at 1 M and 5 M derived from the Fullprof
renement of X-ray diffraction data.
Formula

GdBO3:Eu3
CNaOH 1 M

CNaOH 5 M

Crystal style
Space group
a()
c()
V(3)
Rwp(%)
Rp(%)

Hexagonal
P63/mmc
3.84091
8.92028
113.966
13.3
10.9
8.63

Hexagonal
P63/mmc
3.83791
8.91948
113.779
13.0
9.67
9.34

c2

851

gadolinium atoms in the structure were located on a general position with a bicapped trigonal prismatic coordination as shown in
Fig. 4(c). The capping oxygen atoms were above the triangular
faces, but away from the pseudo-3-fold axis. The gadolinium atoms
were in a C1 symmetry coordination. For the Eu3-doped materials,
the Eu3 ions easily substituted for Gd3 ions in GdBO3 host. And
several emission peaks of 5D0e7FJ (J 1, 2, 3, 4) transitions were
expected. This was exactly what observed experimentally in analogous LnBO3:Eu3 structure [40,41].
The CNaOH value of the colloid solution also has a signicant
effect on the size, morphology and microstructure of GdBO3:Eu3
crystals. Fig. 5 illustrates the FE-SEM images of the as-prepared
GdBO3:Eu3 crystals at different CNaOH values. It could be found
in Fig. 5A that the sample prepared at CNaOH 1 M consisted of
large quantity of mono-concave microspheres. The diameter of the
mono-concave microspheres was about 7e8 mm and their surfaces
were smooth. In order to inspect the microstructure of the monoconcave microspheres better, the corresponding high magnication FE-SEM was carried out and the typical image is shown in
Fig. 5B. It could be clearly found that the mono-concave microsphere was self-assembled by a lot of closely aligned nanoakes
around 20e30 nm in thickness radiating from the center. These
nanoakes were curled to form microsphere, which also led to the
formation of the mono-concave in the center of the microsphere. As
the CNaOH value was gradually increased to 2.0, 3.0 and 4.0, a series
of microowers with near sizes and microstructure are observed in
Fig. 5EeH, respectively. From the corresponding higher magnication SEM images D, F, H, it could be seen that the microowers
were mainly consist of six groups of petal-like nanoakes with
smooth surfaces and uniform thickness. Obviously, most of the
petal-like nanoakes were linked together by both edge-to-edge
and surface-to-surface conjunctions, the petal-like nanoakes
extended outward from the center of the microstructure, and a few
were attached to each other. The numbers of each group petal-like
nanoakes varied from dozens of layers to more than a dozen
layers, consequently to several layers with the CNaOH value gradually increased. Moreover, the sizes of the microowers got bigger
along with the more bloomy of the petal-like nanoakes. Further
increasing the CNaOH value to 5.0, the monodispersed irregular
shaped nanosheets appeared instead of the microowers shown in
Fig. 5I, J. By contrast, the morphology of GdBO3:Eu3 crystals
attained by one step method and solid state reaction are shown in
Fig. S3. It could be observed in Fig. S3(a) and (b) that the products
obtained by one step method were composed of a large number of
nanoparticles and the nanoparticles were uniform in size with an
average diameter of about 40e50 nm. In addition, as presented in
Fig. S3(c) and (d), the morphology of GdBO3:Eu3 crystals synthesized by solid state reaction consisted of plenty of irregular particles
and with smooth surface. The dispersibility of the particles was not
bad and the sizes of the particles were in the range of 3e5 mm.
3.3. Formation mechanism
In order to reveal the morphological evolution of the monoconcave microspheres, diversity microowers and nanosheets
GdBO3:Eu3 crystals, a series of time-dependent control experiments of mono-concave microspheres GdBO3:Eu3 crystals were
carefully and systematically carried out to gain deeper insight into
the formation process. The XRD patterns and the corresponding
SEM images of the mono-concave microspheres GdBO3 intermediates obtained at different reaction time intervals are shown
in Figs. 6 and 7, respectively. They revealed that the intermediates
displayed distinctively different XRD patterns and SEM images at
different reaction periods. After treated the Gd(OH)3 precursors
with H3BO3 for 3 h by the hydrothermal process, the product was

852

Y. Zhu et al. / Journal of Alloys and Compounds 685 (2016) 848e859

Table 2
All atom positions, occupation probability and thermal vibration parameters of GdBO3:Eu3 sample using precursors Gd0.99Eu0.01(OH)3 under CNaOH value at 1 M.
Atom

Wyckoff position

Bios.

Occ.

Gd1
B1
O1
O2

2a
6h
6h
4f

0.00000(0)
0.23200(0)
0.02900(0)
0.33330(0)

0.00000(0)
0.46400(0)
0.05800(0)
0.66670(0)

0.00000(0)
0.25000(0)
0.25000(0)
0.11900(0)

5.626(5)
12.180(4)
0.000(0)
0.000(0)

0.068(2)
0.083(0)
0.033(1)
0.167(0)

Fig. 3. FT-IR spectra of the as-synthesized GdBO3:Eu3 samples when CNaOH 1 M.

mainly Gd(OH)3. A corresponding typical SEM image revealed that


the product was composed of a large number of nanorods. The
diameter and length of the nanorods were less than 50 nm and
range from 200 nm to less than 1 mm. In comparison to the primitive Gd(OH)3 precursors, these intermediates became shorter and
narrower with more smooth surface. As the reaction proceeded up
to 9 h, the phase of GdBO3 appeared, with Gd(OH)3 phases

coexisted this moment. Meanwhile, a number of irregular nanoakes were generated in the intermediate products along with the
preceding nanorods. It was speculated that the irregular nanoakes
was the crystallized GdBO3 crystals from the solution. When the
reaction time reached 12 h, the phase of Gd(OH)3 gradually reduced
and the nal product phase of GdBO3 gradually grew which are
presented in the corresponding XRD patterns. From the corresponding SEM images, we could found that with the increase of
reaction time, some nanoakes grew in a homocentric layer-bylayer and edge-to-edge growth style, to assemble into multilayered disk-like hierarchical structures at the expense of the irregular
nanoakes. And a few nanorods and irregular nanoakes still
remained in the other space. With a reaction time of 18 h, GdBO3
phase further increased with the decreased of precursor Gd(OH)3
phase. At the same time, the nanoakes of the initial formed disklike shape extended at an angle to the plane from the center to form
primary microowers, which was for reducing the surface energy.
Finally, when the reaction time was prolonged to 24 h, GdBO3 phase
was the exclusive product. Simultaneously, the aggregated primary
microowers still underwent the Ostwald ripening process
[42e44] at the expense of the smaller nanoakes to form microsphere shapes. It was also noticed that a concave was presented on
the top of the particles constructed with several curved nanoakes.
On the basis of above experimental results, such a process was
consistent with the previous reports of a two-stage growth process
[28,45,46]. We believed that the formation of GdBO3:Eu3 microsphere, microowers and irregular nanoakes could be rationally
expressed as a kinetically controlled dissolution-recrystallization
mechanism [47e50], which involved a fast dissolution process of

Fig. 4. Crystal structure schematic diagram of GdBO3 (a); the B3O9


9 sheet in the GdBO3 structure (b); coordination of gadolinium atom in the GdBO3 structure (c).

Y. Zhu et al. / Journal of Alloys and Compounds 685 (2016) 848e859

853

Fig. 5. FE-SEM images of the GdBO3:Eu3 phosphors synthesized by Gd0.99Eu0.01(OH)3 precursors with the CNaOH value at (A and B) 1 M; (C and D) 2 M; (E and F) 3 M; (G and H)
4 M; (I and J) 5 M.

the precursors and a slow renucleation and self-assembly process


of the new products. The whole process might be described with
the following equations:

H3 BO3 H2 O/BOH4  H

(1)

GdOH3 H /Gd3 H2 O

(2)

Gd3 BOH4  4H /GdBO3 4H2 O

(3)

The possible conversion process might be as follows: At the


early stage of the reaction, the Gd0.99Eu0.01(OH)3 nanorods started
to dissolve into the solution under the hydrolyzation of H3BO3, and
generated Gd3 ions during the hydrothermal condition. At the
same time, the nanorods became shorter and narrower and with
more smooth surface. With the increase of reaction time, Gd3 ions

steadily increased and B(OH)


could react with Gd3 to
4, H
recrystallize the hexagonal phase of GdBO3. At the moment, the
GdBO3:Eu3 crystals separated out in the solution with irregular
nanoakes. How did the nanoakes form? Mainly due to the

854

Y. Zhu et al. / Journal of Alloys and Compounds 685 (2016) 848e859

dissolution rate of the Gd0.99Eu0.01(OH)3 precursors occurred. These


might result in the lower degree of self-assembly and aggregation,
eventually led to the diversity microowers and nanoakes. These
self-assemblied microspheres and microowers possessed uniform
sizes and excellent dispersion, which were advantages when suspensions were used for fabricating devices where coating uniformity was important. The mechanism of formation of the nal
structure morphologies between primary precursors and products
remained a mystery to materials chemists, although many kinds of
spheres and owerlike 3D structures have been reported. Since
several factors, including crystal-face attraction, electrostatic and
dipolar elds associated with the aggregate, van der Waals forces,
hydrophobic interactions and hydrogen bonds, might have various
effects on the self-assembly process [53,54].
3.4. Effect of PEG on the morphology of GdBO3 crystals
Fig. 6. The XRD patterns of intermediate samples during the synthesis of GdBO3 intermediates collected at different time intervals (3 h, 9 h, 12 h, 18 h, 24 h).

anisotropic crystal structure, there was a tendency for the nucleation growth along the (002) facet which could be veried in Fig. 1
[51,52]. With the reaction proceeding, the nanorods gradually
decreased and transformed into nanoakes as H3BO3 was superuous. Once the concentration of nanoakes reached a point of
supersaturation, the nanoakes started to aggregate into small
clusters via self-assembly. Subsequently, the smaller selfassemblied nanostructures gradually appeared and developed
into multilayered disk-like hierarchical structures. Eventually, the
multilayered structures still underwent the Ostwald ripening process and formed mono-concave microsphere by consuming Gd3
and BO3
3 , which resulted from the dissolution process and hydrolyzed H3BO3. And the concaves were formed due to the
arrangement of the curved nanoakes in revolved order. A schematic illustration of the possible detailed conversion processes
from the precursors to the nal products is presented in Scheme 1.
The formation mechanism of the several microowers and
nanoakes were closely resemblant to the formation mechanism of
the mono-concave microspheres: dissolution-recrystallization
mechanism. Dissimilarly, due to the different basicity of processing precursor solution, the higher basicity i. e. CNaOH, the lower
interfacial energy of the nanostructures surface and slower

PEG with different molecular weights have been successfully


used to prepare various morphologies, such as nanorods and
nanowires [55,56], nanospheres [57], nanocubes and nanoboxes
[58], three-dimensional architectures [59], etc. PEG had hydrophilic
eOe and hydrophobic eCH2eCH2 on the long chains. It could bond
on the surface of rstborn crystal nucleuses because of the coordination action of oxygen atoms, which limited the crystal growth,
and acted as structure-directing reagent as well. The XRD patterns
of the GdBO3 crystals by treating precursors Gd0.99Eu0.01(OH)3
nanorods with different weight percent of PEG (20000) are shown
in Fig. S4. As could be seen from the XRD patterns, all of reections
of the as-synthesized products in the absence of PEG (20000) could
be well indexed to pure hexagonal vaterite phase of GdBO3 (ICSD27934, space group P63/mmc) as set forth. Distinctively, the preferential orientation of (002) facets fade away with the increase of
PEG (20000), which hinted the different morphologies of the
products according to the previous descriptions as Figs. 1 and 2. In
order to prove the above conjecture, the corresponding SEM images
of the GdBO3:Eu3 crystals are exhibited in Fig. 8. Fig. 8A presented
the SEM images of GdBO3:Eu3 crystals obtained when CNaOH value
was 5.0 and in the absence of PEG (20000), the products exhibited
irregular nanosheets morphology just like previous contents.
Fig. 8CeJ showed the morphologies of GdBO3:Eu3 crystals by
processing precursors Gd0.99Eu0.01(OH)3 as CNaOH value was 5.0
with different contents of PEG (20000). When the additive amount
of PEG (20000) was 5 wt% relative to precursors, the morphologies

Fig. 7. SEM images of intermediate samples during the synthesis of GdBO3:Eu3 using Gd0.99Eu0.01(OH)3 as precursors collected at different time intervals: (a) 0 h; (b) 3 h; (c) 9 h;
(d) 12 h; (e) 18 h; (f) 24 h.

Y. Zhu et al. / Journal of Alloys and Compounds 685 (2016) 848e859

855

Scheme 1. Schematic illustration of the possible detailed growth and assembly mechanism of the mono-concave microsphere GdBO3 samples.

of GdBO3:Eu3 turned into monodispersed microowers shown in


Fig. 8C, D, which were similar to the aforementioned microowers
except the discrepancy of the number of composed nanoakes and
curving angle of the petal-like nanoakes. The morphologies of
GdBO3:Eu3 as PEG (20000) dosage was 10 wt% further changed
into mono-concave microspheres seen from Fig. 8E, F with smaller
concave size compared to the previous mono-concave microspheres when CNaOH value was 1.0. As the dosage of PEG (20000)
extended to 15 wt%, completed microdisks composed by several
layers appeared presented in Fig. 8G, H. It could be seen from Fig. 8I,
J that continued to augment the contents of PEG (20000) to 20 wt%,
the morphologies of GdBO3:Eu3 showed mono-concave microspheres again. As can be seen above, as-synthesized products took
on different appearances from irregular nanosheets to completed
microdisks eventually to mono-concave microspheres in the presence of different amounts of PEG (20000). It was chiey because
nanostructures could adsorb PEG (20000) molecules in this hydrothermal reaction. The adsorption of an organic layer eCH2eCH2
around the nanoparticles promoted an (electro) steric stabilization
offsetting the van der Waals interaction that drove the aggregation.
Beyond the sole stabilization, through adsorption driven by the
minimization of the overall free energy of the system, such specic
ligands could enable the ne-tuning of the nanostructures interfacial self-assembly. Such tailored interfacial activity could be advantageously used in the eld of polymer surface modications or
functional coatings via the formation of nanoparticles or multilayers at the surface of a given substrate, supramolecular assembly
[19,60]. Scheme 2 presents the possible procedure of PEG (20000)
effect on the formation from the precursors to the GdBO3:Eu3
crystals.
The progress from turning nanoakes to microspheres through
changing the additive amounts of PEG (20000) happened to be the
inverse process of increasing the CNaOH which realized the transformation from microsphere to nanoakes. It was likely that the
surface interfacial energy of nanostructures might be modied by
basicity of the colloidal solution or the adsorption of PEG (20000)
molecule.
3.5. Luminescent properties
It is well-known that LnBO3 is very excellent and efcient host
lattices for other lanthanide ions, which is appealing to the applications such as optic apparatuses and biological labeling. The
photoluminescence properties of nano-/micro-crystals doped with
lanthanide ions depend not only on their compositions, but also on
their structures, phases, crystallinities, morphologies and sizes.
Hence, in this work, Eu3 was selected as the doping ion to investigate the photoluminescence properties of the GdBO3 with

different morphologies. Fig. 9 exhibits the photoluminescence


excitation and emission spectra of as-synthesized GdBO3:Eu3
samples produced via the solid state reaction and hydrothermal
methods, respectively. There appeared a strong band located at
around 260 nm in the two excitation spectra. This band was
attributed to O2Eu3 charge transfer band (CTB) from the 2p
orbital of O2 to the 4f orbital of Eu3 [61]. The sharp and intense
peaks located in the range from 274 to 313 nm corresponded to the
8
S7/2 / 6IJ and 8S7/2 / 6PJ transitions of Gd3 ions [62,63]. This
conrmed that the energy transfer from Gd3 to Eu3 took place
efciently. The sharp peaks measured from 350 to 475 nm were
attributed to the f-f transitions of Eu3 ions. Among, the observed ff transition at 393 nm was noted to be an intense one. An obvious
difference between the two methods appeared in the excitation
spectra, the hydrothermal methods presented higher intensity of
energy transfer from Gd3 to Eu3 than the f-f transitions of Eu3,
which was reverse in solid state reaction. From Fig. 9 we could see
the emission spectra of the GdBO3:Eu3 which consisted of four
main peak groups that were associated with the transitions from
the excited 5D0 level to 7FJ (J 1e4) levels of Eu3 ions [64]. Among
them, 5D07F1 transition peaked at 593 nm and 5D07F2 transition
at 611 and 626 nm exhibited the bright orange and red light. Known
hexagonal GdBO3 doped with ions exhibited red or orange-red
luminescence under UV excitation. The characteristic emission
spectra of GdBO3:Eu3 consisted mainly of two transition peaks, a
typical magnetic dipole transition 5D07F1 (orange) and a typical
electric dipole transition 5D07F2 (red). The mixed orange and red
emissions degraded the chromaticity of GdBO3:Eu3. The 5D07F2
transition was hypersensitive to the symmetry of the crystal eld
around Eu3 ions, and the intensity of the related emission band
was high if the symmetry was low. In other words, when the Eu3
ions were located in the inversion sites, the 5D07F1 transition was
allowed, while the 5D07F2 transition was forbidden, resulting in
the more intense emission of orange light. The ratio of red to orange
emission intensities, R/O, i.e. I (5D07F2)/I (5D07F1), known as
asymmetric ratio, have been calculated by considering the sum of
the intensities of red emission peaks observed at 611 nm and
626 nm due to the contribution of 5D07F2, divided by the intensity
of the orange peak at 593 nm due to 5D07F1 transition [65]. On the
other hand, photoluminescence intensity ratios of 5D07F2 to
5
D07F1 transitions were often used to study the site symmetry of
Eu3 ions. Compared the obtained R/O values of the two methods, it
could be found that the R/O value of hydrothermal method was
1.3103 relative higher to 0.9458 gotten by solid state reaction. The
enhanced R/O ratio of Eu3 in hydrothermal method implied the
increase of those local environments with reduced symmetries
than solid state reaction. These lower symmetrical environments
were probably induced by lots of surface positions of the

856

Y. Zhu et al. / Journal of Alloys and Compounds 685 (2016) 848e859

Fig. 8. SEM images of the nal products GdBO3:Eu3 with the amount of PEG (20000) at (A and B) 0 wt%; (C and D) 5 wt%; (E and F) 10 wt%; (G and H) 15 wt%; (I and J) 20 wt%.

nanoscaled substructures in hydrothermal method sample [66].


In this experiment, the effect of CNaOH and PEG amounts on the
emission intensity and color purity of the obtained samples with
different morphologies were also investigated. Fig. 10 exhibits the
emission spectra of the GdBO3:Eu3 phosphors obtained by using
precursors Gd0.99Eu0.01(OH)3 under different CNaOH. Moreover, by
analyzing the emission spectra, the relative intensity of 5D07F1
emission and R/O values could be observed in Fig. 11(a). Combined
with aforementioned corresponding SEM images, it could be seen
that the emission intensity of sample could reach up to the
maximum when CNaOH 2 M with microowers morphology

composed by dozens of layers. Why did this microowers shape


own the brightest light contrast to other mono-concave microsphere and microowers morphologies? The reason might be that
this microowers morphology composed by dozens of layers
possessed the biggest specic surface area and approximate
spherelike shape which could efciently absorb the ultraviolet and
emit the light. The R/O value of samples reached maximum when
CNaOH 5 M with nanoakes morphology. It was possibly because
of the increase of local environments around nanoakes with
reduced symmetries than microowers and microsphere. The
preparation-dependent R/O value was attributed to variations in

Y. Zhu et al. / Journal of Alloys and Compounds 685 (2016) 848e859

857

Scheme 2. Schematic illustration of the effect of the PEG (20000) on the formation process of GdBO3 samples.

Fig. 9. The photoluminescence excitation and emission spectra of the as-synthesized


GdBO3:Eu3 samples produced via the solid state reaction and hydrothermal
methods, respectively.

crystalline order, which inuenced the site symmetry near the


Eu3. The above conclusion could be further demonstrated by the
inuence of PEG amount on the emission intensity and R/O values.
Fig. 11(b) summarizes the relative intensity and R/O values for
samples at different amounts of PEG. Likewise, the samples with
microowers and nanoakes morphologies produced the most
intense light and the highest R/O value. The emission intensity and
R/O values of samples still could be changed through varying the
excitation wavelength and so on.

4. Conclusions
In conclusion, we have demonstrated a facile and effective hydrothermal route to synthesize novel architectures (mono-concave
microsphere, microdisk, microowers and nanosheet) of
hexagonal-vaterite types of GdBO3:Eu3 crystals by using

Fig. 10. The emission spectra of the GdBO3:Eu3 phosphors obtained by using precursors Gd0.99Eu0.01(OH)3 under different CNaOH (CNaOH 1.0 M, 2.0 M, 3.0 M, 4.0 M,
5.0 M).

Gd0.99Eu0.01(OH)3 nanorods as precursors. The morphologies of the


obtained architectures could be transformed from mono-concave
microsphere to diversity microowers and then to nanosheet by
treating Gd0.99Eu0.01(OH)3 precursors with different CNaOH. And the
XRD diffraction peaks of (002) facets became stronger with the
CNaOH increased, indicating an obvious preferential growth of
GdBO3:Eu3 crystals. On the contrary, the nal morphologies of
GdBO3:Eu3 could be modulated from nanosheet to microdisk
nally to mono-concave microsphere by adding different amounts
of PEG (20000). Above two factors had inverse effect on the formation of the nal structures, perhaps through altering the interfacial energy of the nanoakes. The reaction mechanism has been
considered as a two-stage growth process, which involved a
dissolution of primary Gd0.99Eu0.01(OH)3 nanorods followed by a
slow crystallization and self-assembly process of GdBO3:Eu3
crystals. It has been proposed that the self-assembly evolved in a
homocentric layer-by-layer growth style. The possible mechanisms

858

Y. Zhu et al. / Journal of Alloys and Compounds 685 (2016) 848e859

Appendix A. Supplementary data


Supplementary data related to this article can be found at http://
dx.doi.org/10.1016/j.jallcom.2016.06.215.
References

Fig. 11. The relative intensity of 5D07F1 emission and R/O values of the GdBO3:Eu3
phosphors obtained by using precursors Gd0.99Eu0.01(OH)3 under CNaOH value at 1 M,
2 M, 3 M, 4 M and 5 M (a); and with the amount of PEG (20000) at 0 wt%, 5 wt%, 10 wt
%, 15 wt% and 20 wt% (b).

have been discussed in detail. Furthermore, this method could be


extended to fabricate other rare-earth orthoborates, and their
morphologies could also be controlled effectively. The as-obtained
Eu3 doped GdBO3 samples showed strong characteristic orange
and red emission of Eu3 under ultraviolet excitation. Contrast to
GdBO3:Eu3 samples synthesized by solid state reaction, the
GdBO3:Eu3 obtained by hydrothermal method possessed much
higher R/O ratio attributed to 5D07F1 and 5D07F2 transitions. The
inuence of diversity morphologies on the R/O ratio and emission
intensity has been also investigated. Results showed that the
microowers structure exhibited the strongest red emission, while
the nanosheets structure emerged the highest R/O ratio. All these
properties indicate that the as-formed GdBO3:Eu3 phosphors are
potentially applied in uorescent lamps and eld emission displays.
More importantly, this general and facile route may be of great
signicance in the synthesis of many other lanthanide compounds
with various shapes.
Acknowledgements
This work was supported by the National Natural Science
Foundation of China (Grant No. 21571162), the Guangdong Province
Enterprise-University-Academy
Collaborative
Project
(No.
2012B091100474) and Open Foundation of Hubei key laboratory of
low dimensional optoelectronic materials and devices (No.
HLOM142002).

[1] O.D. Velev, Self-assembly of unusual nanoparticle crystals, Science 312 (2006)
376e377.
[2] B. Wei, M. Melli, N. Caselli, F. Riboli, D.S. Wiersma, M. Staffaroni, H. Choo,
D.F. Ogletree, S.J. Aloni, B.S. Cabrini, F. Intonti, M.B. Salmeron, E. Yablonovitch,
P.J. Schuck, A. Weber-Bargioni, Mapping local charge recombination heterogeneity by multidimensional nanospectroscopic imaging, Science 338 (2012)
1317e1321.
[3] Y.J. Kang, J.B. Pyo, X.C. Ye, T.R. Gordon, C.B. Murray, Synthesis, shape control,
and methanol electro-oxidation properties of Pt Zn alloy and Pt3Zn Intermetallic nanocrystals, ACS Nano 6 (2012) 5642e5647.
[4] D.Y. Wang, Y.J. Kang, X.C. Ye, C.B. Murray, Mineralizer-assisted shape-control
of rare earth oxide nanoplates, Chem. Mater. 26 (2014) 6328e6332.
[5] Z.G. Yan, C.H. Yan, Controlled synthesis of rare earth nanostructures, J. Mater.
Chem. 18 (2008) 5046e5059.
[6] Y.N. Xia, Y.J. Xiong, B. Lim, S.E. Skrabalak, Shape-controlled synthesis of metal
nanocrystals: simple chemistry meets complex physics? Angew. Chem. Int.
Ed. 48 (2009) 60e103.
[7] M.J. Wang, L.J. Du, X.L. Wu, S.J. Xiong, P.K. Chu, Charged diphenylalanine
nanotubes and controlled hierarchical self-assembly, ACS Nano 5 (2011)
4448e4454.
[8] W.Z. Li, J.D. Fan, J.W. Li, Y.H. Mai, L.D. Wang, Controllable grain Morphology of
Perovskite absorber lm by molecular self-assembly toward efcient solar cell
exceeding 17%, J. Am. Chem. Soc. 137 (2015) 10399e10405.
[9] Y. Zhong, J.F. Wang, R.F. Zhang, W.B. Wei, H.M. Wang, X.P. Lv, F. Bai, H.M. Wu,
R. Haddad, H.Y. Fan, Morphology-controlled self-assembly and synthesis of
photocatalytic nanocrystals, Nano Lett. 14 (2014) 7175e7179.
[10] M. Fialkowski, K.J.M. Bishop, R. Klajn, S.K. Smoukov, C.J. Campbell,
B.A. Grzybowski, Principles and implementations of dissipative (dynamic)
self-assembly, J. Phys. Chem. B 110 (2006) 2482e2496.
[11] G.M. Whitesides, B. Grzybowski, Self-assembly at all scales, Science 295
(2002) 2418e2421.
[12] J.H. Yu, J. Joo, H.M. Park, S.I. Baik, Y.W. Kim, S.C. Kim, T. Hyeon, Synthesis of
quantum-sized cubic ZnS nanorods by the oriented attachment mechanism,
J. Am. Chem. Soc. 127 (2005) 5662e5670.
[13] Z.P. Zhang, H.P. Sun, X.Q. Shao, D.F. Li, H.D. Yu, M.Y. Han, Three-dimensionally
oriented aggregation of a few hundred nanoparticles into monocrystalline
architectures, Adv. Mater. 17 (2005) 42e47.
[14] Z.Y. Tang, Z.L. Zhang, Y. Wang, S.C. Glotzer, N.A. Kotov, Self-assembly of CdTe
nanocrystals into free-oating sheets, Science 314 (2006) 274e278.
[15] M. Remskar, A. Mrzel, Z. Skraba, A. Jesih, M. Ceh, J. Demsar, P. Stadelmann,
vy, D. Mihailovic, Self-assembly of subnanometer-diameter single-wall
F. Le
MoS2 nanotubes, Science 292 (2001) 479e482.
[16] G. Jia, H.P. You, M. Yang, L.H. Zhang, H.J. Zhang, Uniform lanthanide orthoborates LnBO3 (Ln Gd, Nd, Sm, Eu, Tb, and Dy) microplates: general synthesis and luminescence properties, J. Phys. Chem. C 113 (2009)
16638e16644.
[17] C.K. Lin, D.Y. Kong, X.M. Liu, H. Wang, M. Yu, J. Lin, Monodisperse and coreshell-structured SiO2@YBO3:Eu3 spherical particles: synthesis and characterization, Inorg. Chem. 46 (2007) 2674e2681.
[18] J. Ma, Q.S. Wu, Y. Chen, Y. Chen, A synthesis strategy for various pseudovaterite LnBO3 nanosheets via oxides-hydrothermal route, J. Solid State Sci.
12 (2010) 503e508.
[19] Y.F. Xu, D.K. Ma, X.A. Chen, D.P. Yang, S.M. Huang, Bisurfactant-controlled
synthesis of three-dimensional YBO3/Eu3 architectures with tunable wettability, Langmuir 25 (2009) 7103e7108.
[20] M. Ren, J.H. Lin, Y. Dong, L.Q. Yang, M.Z. Su, Structure and phase transition of
GdBO3, Chem. Mater. 11 (1999) 1576e1580.
[21] G. Chadeyron, M. El-Ghozzi, R. Mahiou, A. Arbus, J.C. Cousseins, Revised
structure of the orthoborate YBO3, J. Solid State Chem. 128 (1997) 261e266.
[22] S.S. Liu, D.K. Ma, Y.Q. Zhang, P. Cai, X.A. Chen, S.M. Huang, Controlled synthesis
of orange-like LnBO3: Eu3 (Ln Y, Tb) mesocrystals via a facile organic
additive-free hydrothermal route, CrystEngComm 14 (2012) 2899e2905.
[23] A. Tani, H. Hara, S. Takeshita, T. Isobe, An anomalous downsizing of
glycothermally-synthesized YBO3 crystals by Ce3 doping, CrystEngComm 15
(2013) 10059e10067.
[24] A. Szczeszak, T. Grzyb, B. Barszcz, V. Nagirnyi, A. Kotlov, S. Lis, Hydrothermal
synthesis and structural and spectroscopic properties of the new triclinic form
of GdBO3:Eu3 nanocrystals, Inorg. Chem. 52 (2013) 4934e4940.
[25] G.H. Pan, H.W. Song, X. Bai, Z.X. Liu, H.Q. Yu, W.H. Di, S.W. Li, L.B. Fan, X.G. Ren,
S.Z. Lu, Novel energy-transfer route and enhanced luminescent properties in
YVO4:Eu3/YBO3:Eu3 composite, Chem. Mater. 18 (2006) 4526e4532.
[26] X.C. Jiang, L.D. Sun, C.H. Yan, Ordered nanosheet-based YBO3:Eu3 Assemblies: synthesis and tunable luminescent properties, J. Phys. Chem. B 108
(2004) 3387e3390.
[27] Z.W. Sun, D.M. Liu, L.Z. Tong, J.H. Shi, X.W. Yang, L.X. Yu, Y.C. Tao, H. Yang,
Synthesis and properties of magnetic and luminescent Fe3O4/SiO2/YVO4:Eu3

Y. Zhu et al. / Journal of Alloys and Compounds 685 (2016) 848e859


nanocomposites, Solid State Sci. 13 (2011) 361e365.
[28] Y.B. Zeng, Z.Q. Li, Y.F. Liang, X.Q. Gan, M.M. Zheng, A general approach to
spindle-assembled lanthanide borate nanocrystals and their photoluminescence upon Eu3/Tb3 doping, Inorg. Chem. 52 (2013) 9590e9596.
[29] Z.H. Xu, C.X. Li, Z.Y. Cheng, C.M. Zhang, G.G. Li, C. Peng, Lin, J. Self-assembled
3D architectures of lanthanide orthoborate: hydrothermal synthesis and
luminescence properties, CrystEngComm 12 (2010) 549e557.
[30] Z.H. Leng, N.N. Zhang, Y.L. Liu, L.L. Li, S.C. Gan, Controlled synthesis of different
multilayer architectures of GdBO3:Eu3 phosphors and shape-dependent
luminescence properties, Appl. Surf. Sci. 330 (2015) 270e279.
[31] L. Yang, L.Q. Zhou, Y. Huang, Z.W. Tang, Hydrothermal synthesis of GdBO3:
Eu3 nanobres, Mater. Lett. 64 (2010) 2704e2706.
[32] L.H. Zhang, G. Jia, H.P. You, K. Liu, M. Yang, Y.H. Song, Y.H. Zheng, Y.J. Huang,
N. Guo, H.J. Zhang, Sacricial template method for fabrication of
submicrometer-sized YPO4:Eu3 hierarchical hollow spheres, Inorg. Chem. 49
(2010) 3305e3309.
[33] Z.H. Xu, P.A. Ma, C.X. Li, Z.Y. Hou, X.F. Zhai, S.S. Huang, J. Lin, Monodisperse
coreeshell structured up-conversion Yb(OH)CO3@YbPO4:Er3 hollow spheres
as drug carriers, Biomaterials 32 (2011) 4161e4173.
[34] Z.H. Leng, H.L. Xiong, L.L. Li, N.N. Zhang, Y.L. Liu, S.C. Gan, Facile controlled
synthesis different morphologies of LuBO3:Ln3 (Ln Eu, Tb) phosphors and
tunable luminescent properties, J. Alloys Compd. 646 (2015) 632e638.
[35] G. Jia, Y.H. Zheng, K. Liu, Y.H. Song, H.P. You, Facile surfactant- and templatefree synthesis and luminescent properties of one-dimensional Lu2O3:Eu3
phosphors, J. Phys. Chem. C 113 (2009) 153e158.
[36] Z.H. Xu, C.X. Li, D.M. Yang, W.X. Wang, X.J. Kang, M.M. Shang, J. Lin, Selftemplated and self-assembled synthesis of nano/microstructures of Gd-based
rare-earth compounds: morphology control, magnetic and luminescence
properties, Phys. Chem. Chem. Phys. 12 (2010) 11315e11324.
[37] Y.F. Guo, X. Quan, N. Lu, H.M. Zhao, S. Chen, High photocatalytic capability of
self-assembled nanoporous WO3 with preferential orientation of (002) planes,
Environ. Sci. Technol. 41 (2007) 4422e4427.
[38] F.S. Ke, L. Huang, H.H. Jiang, H.B. Wei, F.Z. Yang, S.G. Sun, Fabrication and
properties of three-dimensional macroporous SneNi alloy electrodes of high
preferential (110) orientation for lithium ion batteries, Electrochem. Commun.
9 (2007) 228e232.
[39] J. Ma, Q.S. Wu, Y.P. Ding, Y. Chen, Assembled synthesis and phase transition of
pseudovaterite NdBO3 layer-by-layer single-crystal nanopancakes via an
oxides-hydrothermal route, Cryst. Growth Des. 7 (2007) 1553e1560.
[40] L. Yang, L.Q. Zhou, X.G. Chen, X.L. Liu, P. Hua, Y. Shi, X.G. Yue, Z.W. Tang,
Y. Huang, Hydrothermal synthesis of YBO3:Tb3 microowers and their
luminescence properties, J. Alloys. Compd. 509 (2011) 3866e3871.
[41] Z.Q. Li, Y.B. Zeng, H.S. Qian, R. Long, Y.J. Xiong, Facile synthesis of GdBO3
spindle assemblies and microdisks as versatile host matrices for lanthanide
doping, CrystEngComm 14 (2012) 3959e3964.
[42] Z.H. Li, J.H. Zeng, Y.D. Li, Solvothermal route to synthesize well-dispersed
YBO3:Eu nanocrystals, Small 3 (2007) 438e443.
[43] D.L. Gao, X.Y. Zhang, W. Gao, Formation of bundle-shaped b-NaYF4 upconversion microtubes via Ostwald ripening, ACS Appl. Mater. Interfaces 5 (2013)
9732e9739.
[44] J. Li, H.C. Zeng, Hollowing Sn-doped TiO2 nanospheres via Ostwald ripening,
J. Am. Chem. Soc. 129 (2007) 15839e15847.
[45] F. Ye, Y. Peng, G.Y. Chen, B. Deng, A.W. Xu, Facile solution synthesis and
characterization of ZnO mesocrystals and ultralong nanowires from layered
basic zinc salt precursor, J. Phys. Chem. C 113 (2009) 10407e10415.
[46] I. Tsuyumoto, M. Kobayashi, T. Are, N. Yamazaki, Nanosized tetragonal BaTiO3
powders synthesized by a new peroxo-precursor decomposition method,
Chem. Mater. 22 (2010) 3015e3020.

859

[47] S. Ding, P. Lu, J.G. Zheng, X.F. Yang, F.L. Zhao, J. Chen, H. Wu, M.M. Wu,
Textured tubular nanoparticle structures: precursor-templated synthesis of
GaN sub-micrometer sized tubes, Adv. Funct. Mater 17 (2007) 1879e1886.
[48] X.L. Hu, Y. Masuda, T. Ohji, K. Kato, Fabrication of Zn(OH)2/ZnO nanosheetZnO nanoarray hybrid structured lms by a dissolutionerecrystallization
route, J. Am. Ceram. Soc. 93 (2010) 881e886.
[49] L. Xu, X.Y. Ji, J.G. Jiang, L. Han, S.N. Che, P. Wu, Intergrown zeolite MWW
polymorphs prepared by the rapid dissolutionrecrystallization route, Chem.
Mater. 27 (2015) 7852e7860.
[50] X.L. Hu, Y. Masuda, T. Ohji, K. Kato, Dissolution-recrystallization induced hierarchical structure in ZnO: bunched roselike and core-shell-like particles,
Cryst. Growth Des. 10 (2010) 626e631.
[51] L.Q. Ye, L. Zan, L.H. Tian, T.Y. Peng, J.J. Zhang, The {001} facets-dependent high
photoactivity of BiOCl nanosheets, Chem. Commun. 47 (2011) 6951e6953.
[52] S.J. Zou, F. Teng, C. Chang, Z.L. Liu, S.R. Wang, Controllable synthesis of uniform
BiOF nanosheets and their improved photocatalytic activity by an exposed
high-energy (002) facet and internal electric eld, RSC Adv. 5 (2015)
88936e88942.
[53] Y. Politi, T. Arad, E. Klein, S. Weiner, L. Addadi, Sea urchin spine calcite forms
via a transient amorphous calcium carbonate phase, Science 306 (2004)
1161e1164.
[54] H. Colfen, M. Antonietti, Mesocrystals: inorganic superstructures made by
highly parallel crystallization and controlled alignment, Angew. Chem. Int. Ed.
44 (2005) 5576e5591.
[55] Y.J. Xiong, Y. Xie, C.Z. Wu, J. Yang, Z.Q. Li, F. Xu, Formation of silver nanowires
through a sandwiched reduction process, Adv. Mater. 15 (2003) 405e408.
[56] X.F. Zhou, D.Y. Zhang, Y. Zhu, Y.Q. Shen, X.F. Guo, W.P. Ding, Y. Chen, Mechanistic investigations of PEG-directed assembly of one-dimensional ZnO
nanostructures, J. Phys. Chem. B 110 (2006) 25734e25739.
[57] T. Waku, M. Matsusaki, T. Kaneko, M. Akashi, PEG brush peptide nanospheres
with stealth properties and chemical functionality, Macromolecules 40 (2007)
6385e6392.
[58] W.Z. Wang, B. Poudel, D.Z. Wang, Z.F. Ren, Synthesis of PbTe nanoboxes using
a solvothermal technique, Adv. Mater. 17 (2005) 2110e2114.
r, L. Nordstierna, M. Andersson, Meso-ordered PEG-based
[59] M. Wallin, A. Altska
particles, Langmuir 31 (2015) 13e16.
[60] L. Qi, J. Fresnais, P. Muller, O. Theodoly, J.F. Berret, J.P. Chapel, Interfacial activity of phosphonated-PEG functionalized cerium oxide nanoparticles,
Langmuir 28 (2012) 11448e11456.
[61] X.Y. Wu, Y.J. Liang, M.F. Zhang, M.H. Tong, D.Y. Yu, Y.L. Zhu, S.Q. Liu, C.J. Yan,
Uniform KCaY(PO4)2:Eu3 phosphors: solegel method, morphology and
luminescence properties, J. Mater Sci. Mater. Electron. 26 (2015) 7324e7330.
[62] M.H. Tong, Y.J. Liang, G.G. Li, Z.G. Xia, M.F. Zhang, F. Yang, Q. Wang, Luminescent properties of single Dy3 ions activated Ca3Gd7(PO4)(SiO4)5O2 phosphor, Opt. Mater. 36 (2014) 1566e1570.
[63] J. He, S. Zhang, J.B. Zhou, J.P. Zhong, H.B. Liang, S.S. Sun, Y. Huang, Y. Tao, Opt.
Mater. 39 (2015) 81e85.
[64] S. Sohala, X. Zhang, V.V. Kuryatkovc, J. Chaudhuri, M. Holtza, Correlation of

photoluminescence and structural order Cin
YBO3:Eu3 micro and nano
structures, Mater. Lett. 106 (2013) 381e384.
[65] U. Rambabu, S.D. Han, Luminescence optimization with superior asymmetric
ratio (red/orange) and color purity of MBO3:Eu3@SiO2 (M Y, Gd and Al)
nano down-conversion phosphors, RSC Adv. 3 (2013) 1368e1379.
[66] Y.P. Li, J.H. Zhang, X. Zhang, Y.S. Luo, S.Z. Lu, X.G. Ren, X.J. Wang, L.D. Sun,
C.H. Yan, Luminescent properties in relation to controllable phase and
morphology of LuBO3:Eu3 nano/microcrystals synthesized by hydrothermal
approach, Chem. Mater. 21 (2009) 468e475.

Você também pode gostar