You are on page 1of 7

Journal of Crystal Growth 438 (2016) 3137

Contents lists available at ScienceDirect

Journal of Crystal Growth


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

Synthesis of hollow spherical tin phosphides (Sn4P3) and their high


adsorptive and electrochemical performance
Shuling Liu a,n, Hongzhe Zhang a, Liqiang Xu b, Lanbing Ma a
a
Key Laboratory of Auxiliary Chemistry & Technology for Chemical Industry, Ministry of Education, Shaanxi University of Science & Technology, Xian,
Shaanxi 710021, PR China
b
School of Chemistry and Chemical Engineering, Shandong University, Jinan 250100, PR China

art ic l e i nf o

a b s t r a c t

Article history:
Received 5 September 2015
Received in revised form
27 November 2015
Accepted 21 December 2015
Available online 29 December 2015

Sn4P3 hollow spherical microstructures are successfully synthesized by a facile Solvothermal approach
with the help of ethylenediamine tetraacetic acid (EDTA). Further study show that EDTA plays an
important role as chelating and capping reagent in the formation of Sn4P3 hollow microspheres at room
temperature. The research displays that the as-prepared Sn4P3 catalysts could absorb some typical
organic dyes such as Methylene Blue, Safranine T and Methyl Orange, especially the adsorption ratio for
some could reach to 99%. Meanwhile, as anode materials for lithium ion batteries, the initial discharge
capacity of the as-prepared Sn4P3 hollow microspheres exceed 1478 mA h g  1 at 100 mA g  1 and it still
keeps at 261mA h g  1 after 20 cycles. Even the current density is increased to 500 mA g  1, the second
discharge capacity still can attain to 689 mA h g  1.
& 2015 Elsevier B.V. All rights reserved.

Keywords:
B1. Sn4P3
A1. Hydrothermal
A2. Hollow
A3. Adsorptive
A4. Electrochemical

1. Introduction
In the last few years, hollow spheres have received much
attention due to their low material density, large surface area,
capacious void and been applied widely in many areas [15]. Until
now, lots of nano/micro functional materials with hollow sphere
structure have been reported, such as metals, alloys, metal oxide,
and transition metal compounds [69].
Transition metal phosphides, a potential classication of
materials, owns particular physical and chemical properties, and
have promising applications in many elds of photochemical catalysis [10], electrochemistry [11,12], photoelectric devices [13,14]
and magnetism [15]. Tin phosphide (Sn4P3), which is one of the
four reported types of tin phosphides [1619], possess a layered
crystalline structure and its special structure makes it having
broad industrial applications, especially in the eld of anode
materials [20] and catalysts [21]. Kim et al. [22] synthesized Sn4P3
nanoparticles using a high energy mechanical milling route, and
showed it having a good cyclability and retained a fairly large
capacity of 370 mA h g  1 up to 50 cycles when cycled within a
limited voltage window. Qian et al. [23] prepared Sn4P3 nanoparticles via a solvtothermal route and a possible mechanism was
proposed under the solvothermal condition. Wu et al. [24] fabricated Sn4P3 lm using a Pulsed Laser Deposition (PLD) method and
n

Corresponding author. Tel.: 86 29 86168315; fax: 86 29 86168312.


E-mail address: shulingliu@aliyun.com (S. Liu).

http://dx.doi.org/10.1016/j.jcrysgro.2015.12.018
0022-0248/& 2015 Elsevier B.V. All rights reserved.

found that the reversible capacity of the Sn4P3 lm maintains


550 mA h g  1 even after 10 cycles. These results have inspired us
to explore the low-temperature synthesis of Sn4P3 nanocrystals
with tailored shapes and investigated their properties. So far,
several synthetic methods have been discovered for the synthesis
of Sn4P3, such as (1) the direct reaction in the quartz tubes with
elemental tin and red phosphorus under higher temperatures
[25]; (2) the chemical vapor method [26]; (3) the hydrothermal/
solvothermal process [27]; (4) the mechanical chemical synthesis
method [28]. Among them, the solvothermal route is one of the
most promising and convenient ones [23], due to its facility and
ease of control. However, the development of a facile, lowtemperature synthesis of such novel shape Sn4P3 nanocrystals
remains a great challenge.
In our experiment, Sn4P3 hollow spheres were reported by a
facile EDTA-assisted solvtothermal route. EDTA was used to control
the growth of the Sn4P3 hollow spheres. Then, the adsorptive
performance for some organic dyes and electrochemical property
as an anode material for lithium ion batteries were both also
studied.

2. Experimental
2.1. Synthesis of Sn4P3 with hollow microspheres
All of the chemical reagents were an analytical pure grade and
used without any purication. In the process of a typical

32

S. Liu et al. / Journal of Crystal Growth 438 (2016) 3137

experiment, an appropriate amount of SnCl2  2H2O (0.452 g,


2 mmol) and NaBH4 (0.074 g, 2 mmol) were dissolved in 40 ml N,
N-dimethylformamide (DMF) to generate a uniform solution. After
vigorously stirring for 30 min at room temperature, the proper
amount of EDTA (0.6 g, 1.8 mmol) was added to the above solution
under stirring, which was then stirred vigorously at room temperature for 30 min, the resultant solution was transferred into a
50 mL Teon-lined stainless steel autoclave. Then, an appropriate
amount of white phosphorus (0.496 g, 4 mmol) was added to the
above system. The autoclave was sealed and maintained at 200 C
for 16 h. Subsequently, it was cooled to room temperature at air.
The resulting black precipitate was ltered and washed successively several times with absolute ethanol, benzene and distilled
water to remove the impurities. Finally, the as-obtained samples
were dried in a vacuum at 60 C for 12 h and collected for
characterization.

polyvinylideneuoride (PVDF). N-Methylpyrrolidone (NMP) was


used as the solvent. The mixed slurry with thickness of 200 m
was coated onto a piece of Cu foil and dried in vacuum oven at
60 C for 12 h, then the resulting foil was roll-pressed and cut into
disks with diameter of 12 mm. Celgard 2300 membrane was used
as the separator. The electrolyte was composed of 1 mol L  1 LiPF6
dissolved ethylene carbonate/dimethyl carbonate/diethyl carbonate (EC/DMC/DEC, volume ratio was 1:1:1). Lithium foils were
used as the counter electrodes. The button batteries were assembled in argon-lled glove box and cycled at the different current
density from 100 to 1000 mA g  1 within voltage limit of 0.013 V.

2.2. Characterization

The as-prepared products were measured by X-ray powder


diffraction. Fig. 1 shows that the typical XRD pattern of the Sn4P3
samples. All of the diffraction peaks can be indexed as the pure
hexagonal Sn4P3 phase with lattice constants of a 3.967 ,
c35.33 , which is almost close to the reported data (JCPDS card
No. 20-1294). The diffraction peaks are intensive and narrow,
which manifest good crystalline nature of the as-synthesized
Sn4P3 microstructures. No characteristic diffraction peaks arising
from other phases or possible impurities, such as Sn, and phosphorus, or any other phosphides, can be detected. To further
conrm the elemental constitution of the as-prepared Sn4P3
microstructures, the EDX measurement was carried out, which is
shown in Fig. 1(b). From Fig. 1(b), it can be easily found that only
the Sn and P peaks are observed, and the Sn/P atomic ratio is 1.416,
which is just close to the stoichiometric ratio of Sn4P3. The
observed EDX peaks and the following Sn/P atomic ratio further
conrm the result of the XRD pattern.
The XPS results (Fig. 2) further prove the above results. As can
be seen from Fig. 2, the as-prepared sample is composed of Sn, P
and O elements and the mole ratio of Sn:P is 1.44:1, which is close
to the atomic ratio of Sn4P3. In addition, one can see that there was
a little oxygen, owing to absorption oxygen in the atmosphere on
the surface of Sn4P3 hollow microspheres.
Fig. 3(a) shows a typical SEM image of the as-prepared Sn4P3
sample. From Fig. 3(a), a large number of spherical structures and
a small number of particles can be easily found. Therein, the
proportion of spherical structure can reach to 70%. The size of
these spherical structures is not very uniform and the diameter
ranges from 1 to 3 mm, while the mean size of particles is about

The X-ray powder diffraction (XRD) patterns of the Sn4P3


samples were measured at room temperature by Bruker D8
advanced X-ray diffractometer equipped with graphitemonochromatized Cu-K radiation ( 1.5418). The transmission electron
microscopy (TEM, JEM-2011) and the Field emission scanning
electron microscope (FESEM; JEOL JSM-6700F) were used to
characterize the morphology and size of the Sn4P3 samples.
2.3. Adsorptive performance
The adsorptive degradation experiments of the organic dyes
were tested for different time durations at room temperature.
30 mg of the Sn4P3 hollow microspheres were ultrasonically dispersed in 50 mL solutions of Safranine T, Methylene Blue and
Methyl Orange aqueous solutions with a concentration of
10 mg L  1, respectively. The mixtures were sonicated for 10 min to
achieve the adsorptiondesorption equilibrium. The optical property changes of the dyes were recorded on a Hach DR5000 UVvis
absorption spectrophotometer with wavelengths ranging from
300 nm to 800 nm.
2.4. Electrochemical measurements
The electrochemical dischargecharge performances of the
samples were tested on a Land battery test system (CT2001A) at
25 C. The working electrodes were consisted of 60 wt% active
materials (Sn4P3 hollow spheres), 30 wt% carbon black, and 10 wt%

3. Results and discussion


3.1. Structure and morphology characterization

Fig. 1. Typical XRD pattern (a) and EDX (b) of as-prepared Sn4P3 sample.

S. Liu et al. / Journal of Crystal Growth 438 (2016) 3137

33

smaller aggregated particles, the size is about 100300 nm, could


also be observed along with their hollow spheres. Moreover, some
as-prepared Sn4P3 hollow microspheres aggregate together, which
also reveals that some small particles easily aggregate into secondary particles, owing to their small dimensions and high surface
energy.
In our experiments, EDTA was found to play a key role in
determining the morphology of the as-prepared Sn4P3 microstructures. As shown from Fig. 4(a), the obtained products were
composed of a large quantity of irregular aggregated particles
which had an average diameter of about 800 nm when no EDTA
was used. As EDTA was increased to 0.35 g, some microspheres
with solid structure (Fig. 4(b)) can be observed. When EDTA was
added to 0.60 g, hollow spheres and particles can be easily found
(Fig. 4(c)). However, further increasing EDTA to 0.80 g, the product
(Fig. 4(d)) was composed of large agglomerates. Thus, we proposed that the amount of EDTA added in the system was
accountable for the formation of Sn4P3 microspheres. It is common
knowledge that EDTA is a good chelating and capping agent [29],
which can adsorb on the surface of Sn4P3 nuclei [30]. When EDTA
is added, a stable Sn-EDTA complex can be formed easily. This
greatly increases the steric hindrance of the elemental Sn, and
helps to control the particles nucleation and growth ratio. When
the amount of EDTA is very low, the capping ability is almost
negligible, and the particles diffused and aggregated together to
form larger spheres, and the capping ability of EDTA became more
and more ascendant with the increase of EDTA, and the nanoparticles were limited to spread. Then, the primary nanoparticles
self-aggregated together to form larger microspheres and thus can
narrow the interfacial energy. While the amount of EDTA is further
increased, the growth of nanostructure is fully suppressed,
therefore, indicating that the chelating effect of EDTA is more
prevailing [31].
3.2. Evaluation of the adsorptive activity of the microcrystals

Fig. 2. XPS analyses of as-prepared Sn4P3 hollow microspheres. (a) Survey; (b) Sn
region; (c) P region.

1 mm. Another, the surface of the microspheres is also relatively


rough and some clear particles can be easily seen, indicating that
the spherical structures are composed of smaller particles. Fig. 3(b)
shows a typical TEM image of as-prepared Sn4P3 sample, the
strong comparison between the dark edge and the center of the
relative brightness provides an testimony of their hollow nature.
The size and wall thickness of most hollow spheres are calculated
to be approximately 1.5 mm and 200 nm, respectively, and the
surface of the microspheres is also not smooth (the inset). Some

As is well known that metal phosphides can be treated as a


good catalyst, which could degraded some organic dyes. Hence
the as-prepared Sn4P3 hollow microspheres were conducted to
adsorb some organic dyes (Methylene Blue, Safranine T and
Methyl Orange). Fig. 5 shows the adsorptive degradation ratios
of three kinds of organic dyes in the existence of different
amounts of Sn4P3 hollow microspheres. From Fig. 5, it can be
indicated that the amount of catalyst has a great effect on the
adsorptive degradation ratio. When the catalyst concentration is
0.1 g L  1, the adsorption rates of the dyes are low, which
Methylene Blue is 57%, Safranine T is 66% and Methyl Orange is
only 2%. With the increase of the catalyst amount (0.3 g L  1),
the degradability is improved greatly. When the amount of
catalyst is further increased to 0.6 g L  1, the adsorption rates of
the dyes are that Methylene Blue is 99%, Safranine T is 88% and
Methyl Orange is 65%. Thus, 0.60 g L  1 is selected as the optimal
concentration.
As shown in Fig. 6(a)(c), the adsorption of three dyes was
further studied in the absence/presence of Sn4P3 hollow microspheres. With the catalyst added, the absorption peaks intensities
of the three dyes weakened signicantly after achieving the
adsorptiondesorption equilibrium when compared to those
without catalyst. That is to say, the catalyst can promote the
adsorption of dyes. Furthermore, it was found that the adsorption
selectivities of the catalyst are also different for the three dyes,
respectively. From Figs. 6 and 7, it can be indicated that, the order
of the adsorption rate of the Sn4P3 hollow microspheres for the
three dyes is Methylene Blue (99%), Safranine T (88%) and Methyl
Orange (65%).

34

S. Liu et al. / Journal of Crystal Growth 438 (2016) 3137

Fig. 3. (a) SEM image of Sn4P3 hollow microspheres, (b) TEM images of Sn4P3 hollow microspheres.

Fig. 4. TEM images of the as-prepared Sn4P3 microstructures with different amounts of EDTA: (a) 0 g; (b) 0.35 g; (c) 0.60 g; (d) 0.80 g.

The surface state of the catalyst plays a key factor which


inuences the adsorptive selectivity and activity. It is well known
that anionic or cationic dye groups are inclined to adsorb on the
surface of catalyst with opposite charges owing to electrostatic
interactions [32], and the electronic structure of the dye is not
changed dramaticlly after the adsorption. In our experiments, the
surface of the Sn4P3 catalyst possess a negative charge (as shown
in the inset of Fig. 7), so it can facilely adsorb dyes with cationic
groups such as Methylene Blue and Safranine T, which makes the
adsorption activities of the Sn4P3 crystals for cationic dyes better
than for Methyl Orange.

3.3. Electrochemical performance


Sn4P3 hollow microspheres are evaluated by a half cell of
Sn4P3/Li. All the electrochemical measurements are carried out
at room temperature between 0.01 and 3.0 V (Vs. Li /Li). Fig. 8
(a) shows the typical rst and second galvanostatic charge
discharge curves of Sn4P3 hollow microspheres electrode at
50 mA g  1. The initial discharge capacity of Sn4P3 hollow
microspheres is as high as 1842.3 mA h g  1, which is much
higher than that of the theoretical capacity, and the second
discharge process of the electrode yields a reversible discharge

S. Liu et al. / Journal of Crystal Growth 438 (2016) 3137

35

capacity around 1047 mA h g  1. The extra initial discharge


capacity of the electrode could be resulted from the formation of
SEI lm during the rst discharging process and the large surface area of Sn4P3 hollow microspheres [2], which provide more
reaction sites on surface. The large quantity of SEI lms help to
consume Li , and cause irreversible insertion/extraction Li . In
addition, there are two plateau regions at 0.750.6 V and 0.25
0.1 V during the rst discharge. The former plateau is probably
due to the reaction between lithium and phosphorus atoms, and
the latter corresponds to the alloying reaction between tin and
lithium [22].
In addition to the high capacity utilization, the Sn4P3 electrode
also demonstrates a considerable rate capability as shown in Fig. 8
(b). All these curves exhibit the similar proles. When the current
density is increased to 100, 200 and 500 mA g  1, the reversible
capacity can still retain 952, 756 and 689 mA h g  1, which still is
higher than the reported value [22]. Even at a current density of
1000 mA g  1, a reversible capacity of 300 mA h g  1 could be
achieved, indicating the excellent rate capability. And the cycling
stability of the Sn4P3 electrode is also investigated at 100 mA g  1.
From Fig. 8(c), it can be seen that the initial discharge capacity is
1480 mA h g  1 and can stay 520 mA h g  1 after 4 cycles. However,
it reduced to 261 mA h g  1 after 20 cycles. The reason may be that
the as-obtained Sn4P3 hollow microspheres are not so stable and
the insertion and extraction of lithium ions leads to the irreversible change of the Sn4P3 hollow microstructures. Fig. 8(d) presents the coulombic efciency of Sn4P3 electrode at 100 mA g  1. It
demonstrates that the Sn4P3 electrode has a low initial coulombic

efciency of 51.3%, which indicates a relatively large irreversible


capacity loss due to the incomplete decomposition of SEI lm
inducing a huge voltage hysteresis and the formation of Li3 P
during the rst cycle. In the subsequent cycles, the coulombic
efciency of the Sn4P3 electrode is gradually steady and almost
near 96%.

Fig. 5. The effect of the catalyst amount on the adsorptive degradation efciency of
Sn4P3 hollow microspheres for different dyes.

Fig. 7. Degradation rates of the Sn4P3 hollow microspheres for different dyes (the
insets are the zeta potential distribution of the catalysts).

4. Conclusion
In this paper, Sn4P3 hollow microspheres were synthesized
successfully via a solvothermal method with EDTA at 200 C for
16 h, it was found that the EDTA acted as a good chelating and
capping reagent in the formation process of the Sn4P3 hollow
microspheres. The adsorptive experiments indicate that the Sn4P3
hollow microspheres can degrade some organic dyes (such as
Methylene Blue, Safranine T and Methyl Orange) to some level,
Moreover, as anode materials for lithium ion batteries, Sn4P3
hollow microspheres reveal that the reversible discharge capacity
around 1047 mA h g  1 at the current density of 50 mA g  1 and a
remarkable rate capability with 66% capacity output at
500 mA g  1 can be achieved, which may be attributed to the
special structures of the as-prepared Sn4P3. So the as-prepared
Sn4P3 hollow microspheres may not only be used as a good
adsorbent, but also still be an alternative negative electrode for
lithium ion batteries.

Fig. 6. UVvis absorption spectra of the different dyes in the absence/presence of Sn4P3 hollow microspheres: (a) Methylene Blue; (b) Safranine T; (c) Methyl Orange.

36

S. Liu et al. / Journal of Crystal Growth 438 (2016) 3137

Fig. 8. Electrochemical performances of the Sn4P3 electrode: (a) initial charge/discharge curves at a constant current of 50 mA g  1; (b) rate performance at various rates
from 50 to 1000 mA g  1; (c) cycle performance at a current rate of 100 mA g  1; and (d) coulombic efciency of Sn4P3 electrode at 100 mA g  1.

Acknowledgments
The authors appreciate the nancial support from the National
Natural Science Foundation of China (21301113 and 21471091), the
Natural Science Foundation of Shaanxi Province of China
(2015JM2059), the Scientic Research Planning Program of the
Education Department of Shaanxi Province (15JK1080), and the
Graduate Innovation Fund of Shaanxi University of Science and
Technology.

References
[1] D.E. Bergbreiter, Self-Assembled, sub-micrometer diameter semipermeable
capsules, Angew. Chem. Int. Ed. 38 (1999) 28702872.
[2] A. Debart, L. Dupont, P. Poizot, J. Leriche, J. Tarascon, A transmission electron
microscopy study of the reactivity mechanism of tailor-made CuO particles
toward lithium, J. Electrochem. Soc. 148 (2001) A1266A1274.
[3] S. Grugeon, S. Laruelle, R. Herrera-Urbina, L. Dupont, P. Poizot, J. Tarascon,
Particle size effects on the electrochemical performance of copper oxides
toward lithium, J. Electrochem. Soc. 148 (2001) A285A292.
[4] C.A. Morris, M.L. Anderson, R.M. Stroud, C.I. Merzbacher, D.R. Rolison, Silica sol
as a nanoglue: exible synthesis of composite aerogels, Science 284 (1999)
622624.
[5] X. Wang, F. Wan, Y. Gao, J. Liu, K. Jiang, Synthesis of high-quality Ni2P hollow
sphere via a template-free surfactant-assisted solvothermal route, J. Cryst.
Growth 310 (2008) 25692574.

[6] J. Li, Y. Ni, K. Liao, J. Hong, Hydrothermal synthesis of Ni12P5 hollow microspheres, characterization and photocatalytic degradation property, J. Colloid
Interface Sci. 332 (2009) 231236.
[7] S. Liu, Y. Qian, L. Xu, Synthesis and characterization of hollow spherical copper
phosphide (Cu3P) nanopowders, Solid State Commun. 149 (2009) 438440.
[8] C.M. Park, K.J. Jeon, Porous structured SnSb/C nanocomposites for Li-ion battery anodes, Chem. Commun. 47 (2011) 21222124.
[9] Y. Zhao, Y. Meng, P. Jiang, Carbon@MnO2 coreshell nanospheres for exible
high-performance supercapacitor electrode materials, J. Power Sources 259
(2014) 219226.
[10] L. Feng, H. Vrubel, M. Bensimon, X. Hu, Easily-prepared dinickel phosphide
(Ni2P) nanoparticles as an efcient and robust electrocatalyst for hydrogen
evolution, Phys. Chem. Chem. Phys. 16 (2014) 59175921.
[11] S. Liu, S. Li, J. Wang, Q. Shi, M. Li, Surfactant-assisted synthesis and electrochemical performances of Cu3P dendrites, Mater. Res. Bull. 47 (2012)
33523356.
[12] S. Liu, L. Yan, H. Li, Solvothermal synthesis of owerlike Co2P nanostructures
and its electrochemical performance, Sci. Adv. Mater. 6 (2014) 746750.
[13] S. Liu, L. Yan, J. Tong, S. Li, H. Li, M. Li, Synthesis of 3D ower-like Cd3P2
microstructures and their optical properties, Micro Nano Lett. 7 (2012)
10191022.
[14] D.D. Lovingood, G.F. Strouse, Microwave induced in-situ active ion etching of
growing InP nanocrystals, Nano Lett. 8 (2008) 33943397.
[15] X. Zheng, S. Yuan, Z. Tian, S. Yin, J. He, K. Liu, L. Liu, Nickel/nickel phosphide
coreshell structured nanoparticles: synthesis, chemical, and magnetic
architecture, Chem. Mater. 21 (2009) 48394845.
[16] P.C. Donohue, Synthesis, structure, and superconducting properties of new
high-pressure forms of tin phosphide, Inorg. Chem. 9 (1970) 335337.
[17] P. Eckerlin, W. Kischio, Darstellung und kristallstruktur der phasen Sn4P3
und Sn4As3, Z. Anorg. Allg. Chem. 363 (1968) 19.
[18] J. Gullman, The crystal structure of SnP, J. Solid State Chem. 87 (1990)
202207.

S. Liu et al. / Journal of Crystal Growth 438 (2016) 3137

[19] O. OLOFSSON, X-ray investigation of the tinphosphorus system, Acta Chem.


Scand. 24 (1970) 11531162.
[20] A. Ueda, M. Nagao, A. Inoue, A. Hayashi, Y. Seino, T. Ota, M. Tatsumisago,
Electrochemical performance of all-solid-state lithium batteries with Sn4P3
negative electrode, J. Power Sources 244 (2013) 597600.
[21] S. Liu, S. Li, M. Li, L. Yan, H. Li, Synthesis of Tin phosphides (Sn4P3) and their
high photocatalytic activities, New J. Chem. 37 (2013) 827833.
[22] Y.U. Kim, C.K. Lee, H.J. Sohn, T. Kang, Reaction mechanism of tin phosphide
anode by mechanochemical method for lithium secondary batteries, J. Electrochem. Soc. 151 (2004) A933A937.
[23] H. Su, Y. Xie, B. Li, X. Liu, Y. Qian, A. Novel One-Step, Solvothermal route to
nanocrystalline Sn4P3, J. Solid State Chem. 146 (1999) 110113.
[24] J.J. Wu, Z.W. Fu, Pulsed-laser-deposited Sn4P3 electrodes for lithium-ion batteries, J. Electrochem. Soc. 156 (2009) A22A26.
[25] L. Hggstrm, J. Gullman, T. Ericsson, R. Wppling, Mssbauer study of tin
phosphides, J. Solid State Chem. 13 (1975) 204207.
[26] R. Binions, C.S. Blackman, C.J. Carmalt, S.A. O'Neill, I.P. Parkin, K. Molloy,
L. Apostilco, Tin phosphide coatings from the atmospheric pressure chemical
vapour deposition of SnX4 (X Cl or Br) and PRxH3  x (R Cyc hex or phenyl),
Polyhedron 21 (2002) 19431947.

37

[27] Y. Xie, H. Su, B. Li, Y. Qian, Solvothermal preparation of tin phosphide nanorods, Mater. Res. Bull. 35 (2000) 675680.
[28] W. Zhou, H. Yang, S. Shao, X. Ai, Y. Cao, Superior high rate capability of tin
phosphide used as high capacity anode for aqueous primary batteries, Electrochem. Commun. 8 (2006) 5559.
[29] Y. Yin, Y. Li, H. Zhang, F. Ren, D. Zhang, W. Feng, L. Shao, K. Li, Y. Liu, Z. Sun,
One-step fabrication of BaMoO4 microstructures with controlled morphologies via a simple EDTA-mediated route, Superlattices Microstruct. 55 (2013)
109117.
[30] S. Liu, H. Liu, J. Zhu, Z. Wang, A complexant-assisted hydrothermal route for
the synthesis of nickel phosphide, J. Mater. Sci. 49 (2014) 75567562.
[31] H. Kim, Je Park, K. Kim, M.K. Han, S.J. Kim, W. Lee, Morphology control of Bi2S3
nanostructures and the formation mechanism, Chin. J. Chem. 31 (2013)
752756.
[32] C. Chen, W. Ma, J. Zhao, Semiconductor-mediated photodegradation of pollutants under visible-light irradiation, Chem. Soc. Rev. 39 (2010) 42064219.

You might also like