You are on page 1of 10

Journal of Colloid and Interface Science 408 (2013) 33–42

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Hydrothermal formation of N-doped (BiO)2CO3 honeycomb-like


microspheres photocatalysts with bismuth citrate and dicyandiamide as
precursors
Qiuyan Li, Haitao Liu, Fan Dong ⇑, Min Fu ⇑
Chongqing Key Laboratory of Catalysis and Functional Organic Molecules, College of Environmental and Biological Engineering, Chongqing Technology and Business University,
Chongqing 400067, PR China

a r t i c l e i n f o a b s t r a c t

Article history: In order to develop efficient visible light driven photocatalysts for environmental application, novel
Received 24 April 2013 N-doped (BiO)2CO3 honeycomb-like hierarchical microspheres were fabricated by an one-pot and tem-
Accepted 18 July 2013 plate-free hydrothermal method, firstly using bismuth citrate and dicyandiamide as precursors. The
Available online 30 July 2013
as-prepared samples were characterized by XRD, SEM, FT-IR, PL, XPS, and UV–vis DRS in detail. The
results indicated that the crystal structure and morphology of the samples can be tuned by hydrothermal
Keywords: reaction temperature. The source of nitrogen doping was from dicyandiamide, which also played a key
Nitrogen doping
role in hydrolyzing bismuth citrate to produce bismuth ions and citrate ions. The ammonium ions from
Hydrothermal synthesis
Growth mechanism
the decomposition of dicyandiamide reacted with bismuth ions and carbonate ions from decomposition
Bismuth subcarbonate of citrate ions, producing in situ N-doped (BiO)2CO3 microspheres. The doped nitrogen substituted for
Visible light photocatalytic oxygen in (BiO)2CO3 and was responsible for the band gap reduction of N-doped (BiO)2CO3. The as-pre-
Hierarchical microspheres pared N-doped (BiO)2CO3 microspheres were applied for removal of NO in air and exhibited excellent vis-
ible light activity, exceeding that of N-doped TiO2 and C-doped TiO2. Time-dependent evolutions of
crystal structure, morphology, chemical composition, and optical property were investigated systemati-
cally to reveal the growth mechanism of the honeycomb-like (BiO)2CO3 microspheres. The growth pro-
cess involved multiple steps, including reaction, nucleation, crystallization, aggregation, and
recrystallization. The proposed growth mechanism could provide new insights into the design and fab-
rication of hierarchical materials with advanced properties.
Ó 2013 Elsevier Inc. All rights reserved.

1. Introduction efficient visible light driven photocatalysts is highly desirable. Re-


cently, bismuth-containing materials with hollow or/and hierar-
Nowadays, environmental pollution and energy crisis have be- chical structure have attracted intensive attention. These
come global issues and attracted intensive attentions. Photocata- bismuth-containing materials include BiOX (X = Cl, Br, I), BiVO4,
lytic technology, which can potentially utilize solar light to BiFeO3, Bi2Ti2O7, Bi2MoO6, Bi2WO6 and Bi2Sn2O7, which exhibited
degrade contaminants and convert solar energy at mild conditions, high photocatalytic performance in degrading organic pollutants
has turned out to be an ideal technology in pollution control and due to their special morphological structure with respect to surface
energy conversion [1]. Recently, photocatalytic materials with areas, pore structure, light harvesting, and charge separation [6–
hierarchical structures were investigated widely due to their en- 14]. The bismuth-containing materials can also be applied in other
hanced properties [2,3]. With the rapid development of nanotech- areas, such as catalysis, drug delivery, energy conversion, and
nology, many novel nanostructured materials with functional nano-devices [2,3,15–17].
properties (optical, electrical, magnetic, catalytic, mechanical, Very recently, (BiO)2CO3 as a new bismuth-containing material
chemical, etc.) were fabricated, which are favorable for the was found to display interesting performance in antibacterial and
advancement of photocatalytic materials [2–5]. environmental applications [18–23]. Compared with other bis-
The traditional photocatalysts (such as TiO2) can only be excited muth-containing materials, (BiO)2CO3 exhibited various micro/
by UV light which only occupies 4% in solar light. In order to nanostructures. Therefore, many researchers were devoted to the
initiate photocatalysts utilize visible light irradiation, developing synthesis and application of (BiO)2CO3 materials. Chen and
co-workers have prepared (BiO)2CO3 nanotubes [18], nanobars
[24], nanoplates [24], cube-like [24], and nanoparticles [19,25].
⇑ Corresponding authors. Fax: +86 23 62769785 605.
Xie’s group have synthesized various (BiO)2CO3 microstructures
E-mail addresses: dfctbu@126.com (F. Dong), fumin1022@126.com (M. Fu).

0021-9797/$ - see front matter Ó 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.jcis.2013.07.040
34 Q. Li et al. / Journal of Colloid and Interface Science 408 (2013) 33–42

for degradation of aqueous dye RhB [20]. (BiO)2CO3 nanosheets and microscopy images. The UV–vis diffuse reflection spectra were ob-
flower-like structures were prepared by Huang et al. [21,22]. Cao’s tained for the dry-pressed disk samples with a Scan UV–vis spec-
group have prepared persimmon-like (BiO)2CO3, exhibiting effi- trophotometer (UV–vis DRS: UV-2450, Shimadzu, Japan), using
cient photocatalytic activity in degrading RhB and eosin sodium BaSO4 as reflectance sample. The Photoluminescence (PL: F-7000,
salt under simulated solar irradiation [23]. Yu and co-workers re- HITACHI, Japan) was used to investigate the optical properties of
ported BiVO4/(BiO)2CO3 nanocompsites achieved by hydrothermal the obtained samples. Nitrogen adsorption–desorption isotherms
treatment which exhibited enhanced visible light photocatalytic were obtained on a nitrogen adsorption apparatus (ASAP 2020,
activity [26]. (BiO)2CO3 hierarchical hollow microspheres were fab- USA) with all samples degassed at 100 °C for 12 h prior to
ricated by Dong’s group through a template-free method and measurements.
exhibited outstanding photocatalytic activity under both UV and
visible light irradiation [27,28]. However, the undoped (BiO)2CO3 2.3. Evaluation of photocatalytic activity
cannot be excited by visible light due to its large band gap. In addi-
tion, the fabrication of N-doped (BiO)2CO3 honeycomb-like hierar- The photocatalytic activity of the as-prepared samples was
chical microspheres using bismuth citrate and dicyandiamide as evaluated by removing NO at ppb level in a continuous flow reac-
precursors has not been reported until now. tor. The volume of the reactor was 4.5 L (30 cm  15 cm  10 cm),
In the present work, we developed a one-pot and template-free made of stainless steel, and covered with Saint-Glass. There was a
hydrothermal method to fabricate N-doped (BiO)2CO3 hierarchical commercial tungsten halogen lamp (150 W) vertically placed out-
microspheres, utilizing bismuth citrate and dicyandiamide as pre- side the reactor. UV cutoff filter (420 nm) was applied to remove
cursors. The as-prepared samples were characterized by various UV light for the test of visible light photocatalytic activity. The
techniques. The morphology and structure of (BiO)2CO3 can be as-prepared sample (0.20 g) was added in distilled water (50 ml)
tuned by hydrothermal reaction temperature. The photocatalytic in a beaker and then coated onto a glass dish with a diameter of
activity test for removal of NO in air was carried out, and the as- 12.0 cm. The coated dish was pretreated at 70 °C to remove water
prepared N-doped (BiO)2CO3 honeycomb-like microspheres exhib- in the suspension and placed in the center of the reactor after cool-
ited excellent visible light photocatalytic activity due to nitrogen ing to room temperature. The NO gas was acquired from a com-
doping and the special morphology. Time-dependent evolutions pressed gas cylinder at a concentration of 100 ppm of NO (N2
of crystal structure, morphology, chemical compositions, and opti- balance, BOC gas) with traceable National Institute of Standards
cal property were investigated systematically to reveal the growth and Technology (NIST) standard. The initial concentration of NO
mechanism of honeycomb-like (BiO)2CO3 microspheres, which was diluted to about 450 ppb by the air stream. The flow rate of
could provide new insights into the design and fabrication of hier- mixed stream of air and NO was controlled at 2.4 L min1 and
archical materials with advanced properties. 15 mL min1, respectively. Then, the two gas streams were pre-
mixed by a three-way valve. After the adsorption–desorption equi-
librium was achieved, the lamp was turned on. The concentration
2. Experimental
of NO was measured by a NOx analyzer (Thermo Scientific, 42i-TL)
every 1 min, which monitors NO and NOx (NOx represents
2.1. Synthesis
NO + NO2). The removal ratio (g) of NO was calculated by g
(%) = (1  C/C0)  100%, C is outlet concentration of NO after reac-
All the reagents employed in this study were analytical grade
tion for time t, and C0 represents the inlet concentration.
and were used without further purification. Distilled water was
used in all experiments. In a typical synthesis process, dicyandia-
mide (0.338 g) was first dissolved in distilled water (75 ml) in a 3. Results and discussion
100 ml autoclave Teflon vessel and stirred for 10 min. Then, bis-
muth citrate (1.60 g) was added to the above solution and further 3.1. Crystal structure
stirred for 30 min. The resulted precursor suspension was hydro-
thermally treated at a certain temperature for 24 h. After that, Fig. 1 shows the XRD patterns of the samples prepared at dif-
the resulting solid was filtered, washed with water and ethanol ferent reaction temperatures. The diffraction peaks of the samples
three times, and dried at 60 °C to get final N-doped (BiO)2CO3 sam- obtained at the temperature of 180 °C and 210 °C can be indexed
ples. In order to investigate the effects of reaction temperature on
the formation of N-doped (BiO)2CO3, the temperature was con-
trolled at 150, 180 and 210 °C and the obtained samples were la-
beled as N-BOC-150 °C, N-BOC-180 °C and N-BOC-210 °C,
respectively. Undoped (BiO)2CO3 sample prepared without adding
dicyandiamide and was labeled as BOC [27]. C-doped TiO2 and N-
doped TiO2 were prepared according to previous report [29,30].

2.2. Characterization

The X-ray diffraction patterns of the samples were collected on


an X-ray diffractometer with Cu Ka radiation (XRD: model D/max
RA, Rigaku Co., Japan). FT-IR spectra were recorded on a Nicolet
Nexus spectrometer over the range of 400–4000 cm1. X-ray pho-
toelectron spectroscopy measurements were carried out to investi-
gate the surface chemical compositions, using Al Ka X-rays
(hm = 1486.6 eV) radiation source operated at 150 W (XPS: Thermo
ESCALAB 250, USA). In order to characterize the morphology of the
obtained products, a scanning electron microscope (SEM: JEOL
model JSM-6490, Japan) was used to collect the scanning electron Fig. 1. XRD patterns of the samples obtained at 150, 180, and 210 °C for 24 h.
Q. Li et al. / Journal of Colloid and Interface Science 408 (2013) 33–42 35

to (BiO)2CO3 (JCPDS-ICDD Card No. 41-1488). Meanwhile, the distribution curves of the samples are shown in Fig. 3b. It can be
intensified (1 1 0) diffraction peak means a preferential orienta- seen that the pore distribution range for N-BOC-180 °C is broad
tion along the (1 1 0) plane. However, when the reaction temper- in the range of 2–50 nm and bimodal with small mesopores
ature is set at 150 °C, the diffraction peaks are obviously different (3.2 nm) and larger ones (22.7 nm), confirming the formation of
from the other two, implying that the sample is not (BiO)2CO3. mesopores. The smaller mesopores can be attributed to the pores
Based on further observation, we can find that the diffraction formed among stacked nanosheets in the microsphere, and the
peaks of N-BOC-150 °C can be assigned to bismuth citrate itself large mesopores can be ascribed to the pores formed among the
[27], which suggests that bismuth citrate cannot be hydrolyzed stack of microspheres. The existence of macropores in N-BOC-
under the temperature of 150 °C. A closer look in Fig. 1 shows 210 °C is further confirmed by the corresponding pore size distri-
that the peak intensity becomes stronger as the reaction temper- bution (Fig. 3b). The pore size distribution is located in the range
ature is increased from 180 °C to 210 °C, which indicates that the of 50–150 nm with a peak at 93.8 nm. Such macropores can be
higher reaction temperature will result in the better crystallinity. ascribed to the pores formed among the randomly aggregated
Note that there is a minor diffraction peak at around 27.5° for N- nanosheets (Fig. 2g). The BET areas of N-BOC-150, N-BOC-180,
BOC-180 °C and N-BOC-210 °C which can be indexed to Bi2O3. and N-BOC-210 °C samples are 4, 18, and 11 m2/g, respectively.
This result suggests that the as-prepared N-doped (BiO)2CO3 sam- The pore volumes of N-BOC-180 and N-BOC-210 °C are 0.106 and
ples consist of minor phase of Bi2O3. According to the crystal 0.069 cm3/g, respectively.
structure of (BiO)2CO3, the (Bi2O2)2+ layers and CO2 3 layers are
orthogonally inter-grown. The internal layered structure would 3.4. Chemical composition by FT-IR and XPS
mediate lower growth rate along certain axis to form nanosheet
morphology [31,32]. FT-IR analysis was carried out to investigate the chemical com-
position of the as-prepared samples, as shown in Fig. 4. As we
3.2. Morphological structure know, the ‘‘free’’ CO23 ions have four internal vibration forms, con-
taining symmetric stretching mode m1, anti-symmetric stretching
The morphology and microstructure of the N-doped (BiO)2CO3 mode m3, out-of-plane bending mode m2, and in-plane bending
samples were characterized by SEM. In Fig. 2, we can easily find mode m4. The peak at 1067 cm1 can be assigned to mode m1. While
that the morphologies of the samples under different reaction tem- the peaks at 846 and 820 cm1 can be attributed to mode m2, the
peratures are obviously different. Accordingly, the reaction tem- mode m3 can be detected at 1468 and 1391 cm1. The peaks at
peratures can affect the morphology of the samples. 698 and 670 cm1 are ascribed to mode m4. The band groups of
Fig. 2a and 2b shows SEM images of N-BOC-150 °C sample. It m1 + m4 at 1756 and 1730 cm1 can also be observed. In addition,
can be seen that the surface of the sample is rough with the peak intensity of CO2 3 ions increases with the increasing reac-
many small particles attached. According to the XRD result, N- tion temperature. The broad peak at 1580 cm1 and 3500 cm1 can
BOC-150 °C is actually bismuth citrate. The rough surface of the be assigned to the vibrations of surface hydroxyl groups and the
N-BOC-150 °C indicates that the reaction has taken place on the adsorbed water molecules [32,33]. Fig. 4 shows that the peak
surface of bismuth citrate. When the temperature reaches 180 °C, intensity at 1580 cm1 decreases as the reaction temperature is in-
large amount of honeycomb-like hierarchical microspheres assem- creased from 180 to 210 °C. As the peak at 1580 cm1 can be as-
bled by nanosheets with an average diameter of 3–4 lm are pro- signed to the vibrations of hydroxyl groups from the adsorbed
duced as shown in Fig. 2c. Obviously, a large number of pores are water molecules, the decrease in hydroxyl groups with increased
formed during the hydrothermal process due to the self-assembly reaction temperature can be attributed to the enhanced crystallin-
of nanosheets. Moreover, the surface becomes smooth compared ity of the samples. As the crystallinity of the sample is enhanced,
with N-BOC-150 °C sample (Fig. 2d). No other morphologies can the adsorbed water molecules are decreased, leading to the de-
be detected, indicating a 100% yield of (BiO)2CO3 superstructure. creased hydroxyl groups. The band at 548 cm1 can be attributed
When the temperature is elevated to 210 °C, another type of mor- to the Bi-O stretching. The bands of N-BOC-150 °C are similar to
phology (loose aggregates) is generated as shown in Fig. 2g and 2 h. bismuth citrate, which is consistent with XRD.
The hierarchical architectures are partly disappeared. The high- The XPS measurement was used to confirm the chemical state
magnification SEM images (Fig. 2b, d and h) reveal that the nano- of the elements on the surface of N-BOC-180 °C and N-BOC-
sheets become thicker as the temperature increased. 210 °C samples, as shown in Fig. 5. The two strong peaks at
The morphological structures of the samples obtained at 180 °C 159.1 and 164.4 eV in the spectra (Fig. 5a) are assigned to Bi4f7/2
and 210 °C are further investigated by TEM. Fig. 2e shows the TEM and Bi4f5/2, respectively, which is the feature of Bi3+ in (BiO)2CO3
image of N-BOC-180 °C, and the entire microsphere is composed of [34]. Fig. 5b shows the C1s spectra of (BiO)2CO3. The spectra can
self-assembled nanosheets. Fig. 2i demonstrates that the N-BOC- be fitted to four peaks at 284.78, 285.93, 287.59, and 288.7 eV,
210 °C is composed of randomly aggregated nanosheets. The respectively. The peaks at 284.78, 285.93, and 287.59 eV can be
HRTEM images of a single nanosheet of N-BOC-180 and N-BOC- attributed to adventitious carbon species, while the peak at
210 °C are shown in Fig. 2f and j. The lattice spacing of N-BOC- 288.7 eV can be ascribed to the CO2 3 in (BiO)2CO3. Meanwhile,
180 and N-BOC-210 °C are determined to be 0.27 nm and the O1s spectra (Fig. 5c) can be fitted to three peaks at 529.9,
0.34 nm, respectively, matching the spacing of the (1 1 0) and 530.99, and 532.2 eV, respectively. The peak at 529.9 eV is charac-
(0 0 4) crystal plane of (BiO)2CO3. teristic of Bi-O in (BiO)2CO3 [35], and the other two peaks at 530.99
and 532.2 eV can be assigned to CO2 3 and hydroxyl groups on the
3.3. Specific BET Surface Areas and Pore Structure surface. N1s spectra (Fig. 5d) were recorded. The binding energy
centered at 400 eV indicates that nitrogen substituted for oxygen
Fig. 3 shows the nitrogen adsorption–desorption isotherms and atom and doped into (BiO)2CO3 lattice during the hydrothermal
the corresponding pore size distribution curves of N-BOC-180 °C process. The atomic concentration of doped nitrogen in N-BOC-
and N-BOC-210 °C. According to the BDDT classification, the major- 180 and N-BOC-210 °C is determined to be 2.1% and 0.50%, respec-
ity of physisorption isotherms can be classified into six types. Typ- tively, indicating that the amount of doped nitrogen is decreased as
ically, the N-BOC-180 °C has an isotherm of type IV, indicating the the reaction temperature is increased. This nitrogen doping effect
presence of mesopores. The N-BOC-210 °C has type III curve, which in (BiO)2CO3 is in accordance with the reported N-doped TiO2 [30].
implies the presence of macropores. The corresponding pore size In our case, the only source of nitrogen doping is dicyandiamide.
36 Q. Li et al. / Journal of Colloid and Interface Science 408 (2013) 33–42

Fig. 2. SEM, TEM, and HRTEM images of the samples obtained at 150 °C (a, b), 180 °C (c, d, e, f), and 210 °C (g, h, i, j) for 24 h.

The dicyandiamide plays dual roles in the formation of (BiO)2CO3 electronic states above the valence band edge, indicating the for-
microspheres, one is to hydrolyze bismuth citrate, and the other mation of mid-gap. Such mid-gap can be ascribed to the nitrogen
is to behave as a doped nitrogen source. The DOS of VB is shown doping and make (BiO)2CO3 microspheres absorb visible light.
in Fig. 5e, and the valence band maximum (VBM) is determined Compared to N-BOC-180 °C, the N-BOC-210 °C sample shows little
to be 1.7 eV. Interestingly, the N-BOC-180 °C sample shows new change in VB due to the very low doping content of nitrogen.
Q. Li et al. / Journal of Colloid and Interface Science 408 (2013) 33–42 37

Fig. 3. N2 adsorption–desorption isotherms (a) and pore size distribution curves (b) of N-BOC-180 and N-BOC-210 °C.

Fig. 4. FT-IR spectra of N-BOC-150, N-BOC-180, and N-BOC-210 °C samples in the range of 400–2000 cm1 (a) and 2000–4000 cm1 (b).

3.5. UV–vis DRS and PL UV and visible light region, respectively. The peak at UV region can
be ascribed to the intrinsic band-to-band transition of (BiO)2CO3
Fig. 6a shows the UV–vis DRS spectra of the as-prepared sam- and the peak at visible region is due to the band-to-band transition
ples. Compared with undoped (BiO)2CO3, the absorption edges of of N-doped (BiO)2CO3. Further observation implies that the peak
N-doped (BiO)2CO3 samples prepared at 180 °C and 210 °C both intensity of N-BOC-180 °C sample is the lowest in both UV and vis-
show a red shift. The N-BOC-180 °C sample especially exhibits ible light region. This fact suggests that the recombination rate of
strong visible light absorption and the absorption edge shifts from electron–hole pairs on N-BOC-180 °C is the lowest. Therefore, we
360 nm to 535 nm as the content of doped nitrogen is high (2.1%). can infer that N-BOC-180 °C sample may exhibit the most efficient
However, compared to N-BOC-180 °C, the absorption edge of N- photocatalytic activity.
BOC-210 °C is blue-shifted to 375 nm, which is caused by the lower
amount of doped nitrogen (0.50%). Such shift in absorption edge
can also be reflected by the change in color1 of the samples from 3.6. Photocatalytic activity
yellow to light gray. As shown in Fig. 6b, the band gap energies of
N-BOC-150 °C, N-BOC-180 °C, and N-BOC-210 °C samples are esti- NO is a representative indoor and outdoor air pollutant. The
mated to be 3.16, 2.38, and 3.15 eV, respectively, which are lower photocatalytic activity of the as-prepared samples is evaluated by
than that of undoped (BiO)2CO3 (3.33 eV) [27]. The reduction in removal of NO in gas phase. As previous reported, NO was very sta-
the band gap can be assigned to the in situ doped nitrogen in ble and cannot be photolyzed under light irradiation in the absence
(BiO)2CO3, which makes the samples absorb visible light. This result of photocatalyst [34]. In the presence of the as-prepared samples,
is also consistent with new electronic states observed in VB XPS. NO reacted with the photogenerated reactive radicals to produce
Similar effects have been widely observed in N-doped TiO2 [30]. HNO2 and HNO3 [33,34].
The band gap of N-BOC-180 °C sample is greatly reduced to Fig. 8 shows the variation of NO concentration (C/C0%) with
2.38 eV, which indicates the potential excellent photocatalytic activ- irradiation time over N-BOC-150 °C, N-BOC-180 °C, and N-BOC-
ity under visible light irradiation. 210 °C samples under visible light irradiation with C-doped TiO2
Fig. 7 shows the PL spectra of the N-BOC-150 °C, N-BOC-180 °C, and N-doped TiO2 as references. Obviously, the N-BOC-150 °C sam-
and N-BOC-210 °C samples with an excitation wavelength of ple exhibits negligible visible light activity because N-BOC-150 °C
280 nm. It can be seen that the fluorescence emission peaks are is actually bismuth citrate. The NO removal ratio of C-doped TiO2
mainly centered at 325–400 nm and 450–500 nm, which represent and N-doped TiO2 is 25% and 36%, respectively. However,
N-BOC-180 °C sample shows outstanding visible light photocata-
lytic activity with a NO removal ratio of 41%, much higher than
1
For interpretation of color in Fig. 6, the reader is referred to the web version of that of C-doped TiO2 and N-doped TiO2, although the surface area
this article. of N-BOC-180 °C (18 m2/g) is smaller than that of C-doped TiO2
38 Q. Li et al. / Journal of Colloid and Interface Science 408 (2013) 33–42

Fig. 5. XPS spectra of N-BOC-180 and N-BOC-210 °C, Bi4f (a), C1s (b), O1s (c), N1s (d), and VB (e).

(123 m2/g) and N-doped TiO2 (63 m2/g). This result suggests that light absorption (Fig. 6), and low recombination rate of electron–
N-doped (BiO)2CO3 microspheres are excellent visible light photo- hole pairs (Fig. 7).
catalysts due to its intrinsic property. The hierarchical superstruc-
ture, which belongs to the intrinsic property, is in favor of the 3.7. Growth mechanism of hierarchical microspheres
diffusion of reaction intermediates to accelerate the reaction rate.
The special superstructure also allows multiple reflections of light, In order to investigate how to form such special hierarchical
which enhance the light-harvesting efficiency and increase the nanosheet architectures, time-dependent experiment was carried
quantity of photogenerated electrons and holes to participate in out to understand the detailed evolution process during hydrother-
the photocatalytic reaction, thus enhancing the photocatalytic mal reaction. The products obtained at different reaction times (6, 9,
activity. The NO removal ratio of N-BOC-210 °C (33%) is lower than 12, 16, 24, and 48 h) were analyzed by several different characteriza-
that of N-BOC-180 °C, which can be attributed to the destroyed tions. The reaction temperature was controlled at 180 °C with other
hierarchical structures (Fig. 2g), the small surface areas and the conditions unchanged. The resulted products at different reaction
low amount of doped nitrogen (0.50%). In general, the visible light times were labeled as N-BOC-6 h, N-BOC-9 h, N-BOC-12 h,
photocatalytic activity of N-BOC-180 °C is the best due to the con- N-BOC-16 h, N-BOC-24 h, and N-BOC-48 h, respectively. Time-
tribution of hierarchical superstructure (Fig. 2c), strong visible dependent evolutions of phase structure, chemical composition,
Q. Li et al. / Journal of Colloid and Interface Science 408 (2013) 33–42 39

Fig. 6. UV–vis DRS (a) and the plot of (ahm)1/2 vs. photon energy (b) of N-BOC-150, N-BOC-180, N-BOC-210 °C, and undoped (BiO)2CO3 samples.

diffraction peaks of the samples obtained at every stage are signif-


icantly different, indicating that the microstructures of the samples
grow gradually by prolonging the reaction time. When the reaction
time reaches 6 h, the diffraction peaks of the sample can be in-
dexed to bismuth citrate. Compared with Fig. 1, we can see that
the diffraction peak of N-BOC-6 h is similar to that of N-BOC-
150 °C. Therefore, N-BOC-6 h is determined to be bismuth citrate.
By extending the reaction time to 9 h, the peak intensity becomes
weaker compared with that of N-BOC-6 h, implying that bismuth
citrate starts to be hydrolyzed and intermediate product is formed.
Bismuth citrate is first hydrolyzed by a large amount of OH ions
provided by the dissolved NH3 which comes from the decomposi-
tion of dicyandiamide. The BiO+ and citrate ions are produced
simultaneously (reactions (1) and (2)). Under the hydrothermal
temperature of 180 °C, the citrate ions can be decomposed to pro-
duce carbonate ions (reaction (3)). However, bismuth citrate can-
not be hydrolyzed under the same conditions without
dicyandiamide, suggesting that dicyandiamide plays an important
Fig. 7. PL spectra of N-BOC-150 °C, N-BOC-180 °C, and N-BOC-210 °C samples. role in hydrolyzing bismuth citrate to produce citrate ions.

NH3 þ H2 O ! NHþ4 þ OH ð1Þ

þ
BiðC6 H5 O7 Þ þ 3OH ! BiO þ C6 H5 O3
7 þH
þ
ð2Þ

C6 H5 O3 2
7 ! CO2 þ H2 O ! CO3 þ H
þ
ð3Þ
When the reaction time is prolonged to 12 h, all the diffraction
peaks are almost disappeared, suggesting that most of bismuth cit-
rate has been hydrolyzed. At this stage, there exist a large number
of ions in the solution which react with each other to form semi-
crystallized intermediate. By increasing the reaction time to 16 h,
the diffraction peaks become distinct and can be indexed to the
(BiO)4CO3(OH)2 phase (JCPDS-ICDD Card No. 38-0579), although
the crystallinity is not good. The (BiO)4CO3(OH)2 is produced by
the reaction (4) on the basis of previous reactions.
þ
4BiO þ CO2 
3 þ 2OH ! ðBiOÞ4 CO3 ðOHÞ2 ðSÞ ð4Þ

Fig. 8. Visible light photocatalytic activity of N-BOC-150, N-BOC-180, N-BOC- After reaction for 24 h, the diffraction peaks become stronger
210 °C, C-doped TiO2, and N-doped TiO2 for the removal of NO in air. and can be indexed to (BiO)2CO3 phase (JCPDS-ICDD Card No. 41-
1488). This result indicates that CO2 
3 substitutes for OH gradually
morphological structure, and optical property were investigated in (BiO)4CO3(OH)2 and the phase is transformed to (BiO)2CO3 be-
systematically to reveal the growth mechanism of such novel tween 16 and 24 h (reaction (5)). All the (BiO)4CO3(OH)2 interme-
superstructure. diates are transformed to (BiO)2CO3 phase after reaction for 24 h.
The enlarged view (Fig. 9b) shows that the diffraction angle of
3.7.1. Evolution of phase structure the dominant peaks at around 27–30° shifts slightly to higher val-
Fig. 9a shows the XRD patterns of the samples prepared at ues by increasing the reaction time. This fact suggests that the
different reaction stages from 6 h to 48 h. It is obvious that the phase transition from (BiO)4CO3(OH)2 to (BiO)2CO3 took place
40 Q. Li et al. / Journal of Colloid and Interface Science 408 (2013) 33–42

Fig. 9. XRD patterns of the samples obtained at 180 °C for different reaction times from 6 to 48 h (a), enlarged view of (1 0 3) and (1 1 0) diffraction region (b).

during this reaction process. When the reaction time is prolonged 3.7.3. Evolution of morphological structure
to 48 h, the peak intensity becomes stronger and the pure (BiO)2- SEM analysis was carried out to directly observe the morpho-
CO3 phase with good crystallinity is obtained. During the forma- logical evolution of the samples between 6 h and 48 h. As shown
tion, crystalline ðBiOÞ2 CO3 ; NHþ
4 is simultaneously doped into the in Fig. 11, the morphological structure at every stage is signifi-
lattice of (BiO)2CO3, subsequently modifying its band structure. cantly different. Specifically, the growth process of (BiO)2CO3 hier-
archical superstructures includes the following stages:
ðBiOÞ4 CO3 ðOHÞ2 ðSÞ þ CO2
3 ! ðBiOÞ2 CO3 ðSÞ þ 2OH

ð5Þ Initially, the precursor suspension is composed of bismuth cit-
rate and dicyandiamide. After hydrothermal treatment for 6 h, sub-
stantial irregular particles are obtained as shown in Fig. 11a and b.
3.7.2. Evolution of chemical composition
According to the results of XRD and FT-IR, these irregular particles
In order to investigate the evolution of chemical composition of
are bismuth citrate. When the reaction time reaches 9 h, the SEM
the samples during the hydrothermal process, FT-IR analysis was
images (Fig. 11c and d) obviously show that the microstructures
carried out. The spectra of the samples obtained at different reac-
become loose, indicating that bismuth citrate starts to be hydro-
tion stages are shown in Fig. 10. At first sight, we can find that
lyzed (reaction 2). The XRD pattern of N-BOC-9 h can also prove it.
the bands of the samples obtained at 6 h and 9 h are distinctly dif-
After reaction for 12 h, the microstructures become looser as
ferent from other samples. Based on further observation, the dif-
shown in Fig. 11e. Some semi-crystallized irregular particles cir-
fraction peaks of N-BOC-6 h are similar to that of bismuth citrate,
cled in Fig. 11f are produced during the stage due to the subse-
consistent with the result of XRD. When the reaction time is in-
quent process of reaction, nucleation, aggregation, and
creased to 9 h, the bands of bismuth citrate become weaker. Some
crystallization. A large number of ions came from the hydrolysis
peaks at 846 cm1 and 670 cm1 appeared and can be ascribed to
of bismuth citrate react with each other in the solution to form nu-
CO2
3 . The result implies that bismuth citrate starts to be hydro-
cleus. Then, the nucleus continues to crystallize and small particles
lyzed and CO2 3 ions are gradually formed. From 12 h to 16 h, the
are generated. These small particles dissolve and recrystallize to
peak intensity of CO23 becomes stronger. When the reaction time
form large particles. The consumption of small particles and gener-
is increased to 24 h, the peak intensity of hydroxyl groups in
ation of large particles are complied with the well-known ‘‘Ost-
(BiO)4CO3(OH)2 decreases rapidly and the peak intensity of carbon-
wald ripening’’ mechanism [36,37].
ate ions increases quickly with the prolonged reaction time,
After reaction for 16 h, a large amount of honeycomb-like
which can be ascribed to the CO2 
3 substitutes for OH and makes
microspheres are obtained as shown in Fig. 11g. While they are
(BiO)4CO3(OH)2 transform into (BiO)2CO3. Combining the result of
stacked together and interconnected by the un-dissolved matter,
XRD with FT-IR, it can be concluded that (BiO)4CO3(OH)2 is an
some small particles are still adhered to the surface of the sample
important intermediate on the way to form (BiO)2CO3.

Fig. 10. FT-IR spectra of the samples obtained at 180 °C for different reaction times from 6 to 48 h in the range of 400–2000 cm1 (a) and 2000–4000 cm1 (b).
Q. Li et al. / Journal of Colloid and Interface Science 408 (2013) 33–42 41

Fig. 12. UV–vis DRS spectra of the samples obtained at 180 °C for different reaction
time from 6 to 48 h.

After reaction for 24 h, the honeycomb-like hierarchical micro-


spheres with an average diameter of 3–4 lm are dispersed from
the stacks as shown in Fig. 11i. All the small particles attached to
the surface are disappeared. The enlarged view (Fig. 11j) shows
that the nanosheets become thicker compared with the sample ob-
tained at 16 h, indicating that the nanosheets continue to grow be-
tween 16 h and 24 h. The disappeared small particles take part in
the dispersion and anisotropic growth of the microspheres. Well-
defined (BiO)2CO3 hierarchical microspheres self-assembled by
nanosheets are obtained at this stage. The XRD pattern and FT-IR
spectra of N-BOC-24 h can prove it. When the reaction time is in-
creased to 48 h, the morphological structures have little change,
suggesting that 24 h is an appropriate time to prepare (BiO)2CO3
hierarchical microspheres.

3.7.4. Evolution of optical property


In order to investigate the evolution of optical property of N-
doped (BiO)2CO3, UV–vis DRS analysis was carried out, as shown
in Fig. 12. It can be seen that the samples display strong absorption
in the UV–visible region from 220 to 600 nm. When the reaction
time is 6 h, the absorption edge of the sample is at about
450 nm. By extending the reaction time from 9 to 16 h, the absorp-
tion edge of the samples shows an obvious red shift, indicating that
band gap transition took place during this stage due to nitrogen
doping. Finally, the absorption edge of N-BOC-24 h reaches
550 nm which is much larger than that of undoped (BiO)2CO3.
Interestingly, the spectra of the samples obtained at 9, 12, and
16 h show two absorption edges, suggesting that the samples con-
tain two semiconductors during the reaction process. Combining
the result with XRD and FT-IR, we can find that one of the semicon-
ductors is (BiO)4CO3(OH)2, which has been identified as an impor-
Fig. 11. SEM images of samples obtained at 180 °C for different reaction time, 6 h tant intermediate. When the reaction time reaches 24 h, the
(a, b), 9 h (c, d), 12 h (e, f), 16 h (g, h), 24 h (i, j), and 48 h (k, l).
spectra display only one absorption edge, suggesting that
(BiO)4CO3(OH)2 has completely transformed into N-doped
as circled in Fig. 11g, indicating that the sample is not transformed (BiO)2CO3. When the reaction time is prolonged to 48 h, the visible
into (BiO)2CO3 completely, which can be proved by the XRD pat- light absorption of the sample is weakened, suggesting that 24 h is
tern of N-BOC-16 h. Enlarged view in Fig. 11h shows that the an optimized reaction time.
microstructure of the sample is hierarchical microsphere self-
assembled by many nanosheets. Given such a tremendous change, 4. Conclusion
a series of reactions must take place between 12 h and 16 h (reac-
tion 4 and 5). The production of the nanosheets is at the expense of In summary, we have successfully fabricated novel N-doped
large particles. Then, the nanosheets gradually aggregated together (BiO)2CO3 honeycomb-like hierarchical microspheres. The synthe-
to form microspheres. sis was based on a hydrothermal treatment of bismuth citrate and
42 Q. Li et al. / Journal of Colloid and Interface Science 408 (2013) 33–42

dicyandiamide mixtures. The morphology and photocatalytic [2] X.W. Lou, L.A. Archer, Z.C. Yang, Adv. Mater. 20 (2008) 3987.
[3] Q.J. Xiang, J.G. Yu, M. Jaroniec, Chem. Soc. Rev. 41 (2012) 782.
activity of the samples can be simply tuned by reaction tempera-
[4] Y. Zhao, L. Jiang, Adv. Mater. 21 (2009) 3621.
ture. The precursor dicyandiamide played two important roles in [5] G. Liu, L. Wang, H. Yang, H. Cheng, G.Q. Lu, J. Mater. Chem. 20 (2010) 831.
the formation of N-doped (BiO)2CO3 honeycomb-like hierarchical [6] C. Zhang, Y.F. Zhu, Chem. Mater. 17 (2005) 3537.
microspheres. One is to act as nitrogen doping source and the other [7] F. Amano, K. Nogami, M. Tanaka, B. Ohtani, Langmuir 26 (2010) 7174.
[8] L.W. Zhang, T.G. Xu, X. Zhao, Y.F. Zhu, Appl. Catal. 98 (2010) 138.
is to hydrolyze bismuth citrate. The N-doped (BiO)2CO3 had nar- [9] J.Q. Yu, A. Kudo, Adv. Funct. Mater. 16 (2006) 2163–2169.
rowed band gap and could absorb visible light due to the doped [10] X. Zhang, Z.H. Ai, F.L. Jia, L.Z. Zhang, J. Phys. Chem. 112 (2008) 747.
nitrogen substituting for oxygen in (BiO)2CO3. The as-prepared [11] Y.N. Huo, M. Miao, Y. Zhang, J. Zhu, H.X. Li, Chem. Commun. 47 (2011) 2089.
[12] Z.F. Bian, Y.N. Huo, Y. Zhang, J. Zhu, Y.F. Lu, X. Li, Appl. Catal. 91 (2009) 247.
N-doped (BiO)2CO3 microspheres exhibited excellent visible light [13] J.J. Wu, F.Q. Huang, X.J. Lu, P. Chen, D.Y. Wan, F.F. Xu, J. Mater. Chem. 21 (2011)
photocatalytic activity for NO removal due to the nitrogen doping 3872.
and special hierarchical architecture, even exceeding that of N- [14] S. Shamaila, A.K.L. Sajjad, F. Chen, J.L. Zhang, J. Colloid Interface Sci. 356 (2011)
465.
doped TiO2 and C-doped TiO2. The outstanding photocatalytic [15] L.P. Zhu, H.M. Xiao, W.D. Zhang, G. Yang, S.Y. Fu, Cryst. Growth Des. 8 (2008)
activity of N-doped (BiO)2CO3 microspheres suggests that it is a 957.
very promising visible light driven photocatalyst for environmen- [16] J.G. Yu, H.G. Yu, H.T. Guo, M. Li, S. Mann, Small 4 (2008) 87.
[17] L.S. Zhang, W.Z. Wang, L. Zhou, H.L. Xu, Small 3 (2007) 1618.
tal pollution control. Based on direct time-dependent observations [18] R. Chen, M.H. So, J. Yang, F. Deng, H.Z. Sun, Chem. Commun. (2006) 2265.
of phase structure, chemical composition, morphology and optical [19] R. Chen, G. Cheng, M.H. So, J.L. Wu, Z. Lu, C.M. Che, H.Z. Sun, Mater. Res. Bull. 45
property, a new growth mechanism of honeycomb-like (BiO)2CO3 (2010) 654.
[20] Y. Zheng, F. Duan, M.Q. Chen, Y. Xie, J. Mol. Catal. A: Chem. 317 (2010) 34.
microspheres was revealed. (BiO)4CO3(OH)2 was identified as an
[21] Y.Y. Liu, Z.Y. Wang, B.B. Huang, K.S. Yang, X.Y. Zhang, X.Y. Qin, Y. Dai, Appl.
intermediate emerged during this process. The growth mechanism Surf. Sci. 257 (2010) 172.
involved unique multiple processes, including reaction, nucleation, [22] H.F. Cheng, B.B. Huang, K.S. Yang, Z.Y. Wang, X.Y. Qin, X.Y. Zhang, Y. Dai,
crystallization, aggregation, and recrystallization. The proposed ChemPhysChem 11 (2010) 2167.
[23] X.F. Cao, L. Zhang, X.T. Chen, Z.L. Xue, CrystEngComm 2011 (1939) 13.
growth mechanism could provide new insights into the design [24] G. Cheng, H.M. Yang, K.F. Rong, Z. Lu, X.L. Yu, R. Chen, J. Solid State Chem. 2010
and controlled synthesis of hierarchical materials with enhanced (1878) 183.
properties. [25] G. Cheng, J.L. Wu, F. Xiao, H. Yu, Z. Lu, X.L. Yu, R. Chen, Mater. Lett. 63 (2009)
2239.
[26] P. Madhusudan, J.R. Ran, J. Zhang, J.G. Yu, G. Liu, Appl. Catal. B 110 (2011) 286.
Acknowledgments [27] F. Dong, A.M. Zheng, Y.J. Sun, M. Fu, B.Q. Jiang, W.K. Ho, S.C. Lee, Z.B. Wu,
CrystEngComm 14 (2012) 3534.
[28] F. Dong, W.K. Ho, S.C. Lee, Z.B. Wu, M. Fu, S.C. Zou, Y. Huang, J. Mater. Chem. 21
This research is financially supported by the Science and Tech- (2011) 12428.
nology Project from Chongqing Education Commission (KJ120701, [29] F. Dong, H.Q. Wang, Z.B. Wu, J. Phys. Chem. C 113 (2009) 16717.
KJ130725, [2011]65, KJZH11214), the Natural Science Foundation [30] Z.B. Wu, F. Dong, W.R. Zhao, S. Guo, J. Hazard. Mater. 157 (2008) 57.
[31] C.Z. Wu, Y. Xie, Chem. Commun. (2009) 5943.
Project of CQ CSTC (cstc2013jcyjA20018, cstc2012jjA20014), the [32] F. Dong, S.C. Lee, Z.B. Wu, Y. Huang, M. Fu, W.K. Ho, S.C. Zou, B. Wang, J. Hazard.
National Natural Science Foundation of China (51108487), the Mater. 195 (2011) 346.
Innovative Research Team Development Program in University of [33] F. Dong, Y.J. Sun, M. Fu, W.K. Ho, S.C. Lee, Z.B. Wu, Langmuir 28 (2012) 766.
[34] Z.H. Ai, W.K. Ho, S.C. Lee, L.Z. Zhang, Environ. Sci. Technol. 43 (2009) 4143.
Chongqing (KJTD201314, KJTD201020), and the key discipline
[35] S. Shamaila, A.K.L. Sajjad, F. Chen, J.L. Zhang, Appl. Catal. 272 (2010) 280.
development project of CTBU (1252001). [36] J.B. Lian, X.C. Duan, J.M. Ma, P. Peng, T. Kim, W.J. Zheng, ACS Nano 3 (2009)
3749.
References [37] H.C. Zeng, Curr. Nanosci. 3 (2007) 177.

[1] Y.H. Lan, X.Z. Qian, C.J. Zhao, Z.M. Zhang, X. Chen, Z. Li, J. Colloid Interface Sci.
395 (2013) 75.

You might also like