Sei sulla pagina 1di 12

Journal of Thermal Analysis and Calorimetry (2020) 140:2131–2142

https://doi.org/10.1007/s10973-019-08896-0

Physicochemical and photocatalytic properties of tin dioxide


supported onto silica gel
S. Khalameida1 · V. Sydorchuk1 · S. Levytska1 · N. Shcherban2

Received: 27 March 2019 / Accepted: 6 October 2019 / Published online: 15 November 2019
© Akadémiai Kiadó, Budapest, Hungary 2019

Abstract
The samples containing 1–10% tin dioxide supported onto silica gel have been synthesized via precipitation or thermolysis
of tin tetrachloride. The prepared samples have been characterized using DTA–TG, XRD, Raman and UV–Vis spectroscopy,
TPD of ammonia, and low-temperature nitrogen adsorption–desorption. The supported samples have been tested as photo-
catalysts in the process of rhodamine B degradation under visible irradiation. It has been established that the deposited phase
is uniformly dispersed on the surface. Redshift of band gap is observed for the supported samples. Unlike a bulk ­SnO2, the
supported samples exhibit photocatalytic activity under visible irradiation.

Keywords  Tin dioxide · Silica gel · Photocatalysis · DTA–TG · UV–Vis spectroscopy

Introduction simultaneously optimizes the porosity, creating a meso-


macroporous structure (the most suitable structure for most
Tin dioxide ­SnO2 is a versatile material since it has wide adsorption and catalytic processes) as it was demonstrated
numerous applications in a bulk and deposited state as cata- for different perovskites and ­SnO2 [4, 5]. Another approach
lyst, ion-exchanger, and material for gas sensors [1–3]. Being can be the deposition of an active phase onto the surface
a semiconductor by its nature, it is a promising photocatalyst of mesoporous supports as it was shown for ­TiO2 [8–10],
for processes of organic pollutants degradation [4–6]. The ­BaTiO3 [11], and heteropolycompounds [12, 13]. However,
efficiency of these applications is determined by crystal and this method practically was not investigated for the prepara-
porous structure, surface design, and electronic character- tion of supported photocatalysts based on tin dioxide. The
istics of ­SnO2. The main disadvantages of tin dioxide as a exceptions are the papers on photodegradation of rhodamine
photocatalyst are (i) large band gap (3.4–4.0 eV [1, 4, 6]); B (RhB) under UV irradiation using tin dioxide supported
as a result, this material is active only under UV irradia- on porcine bone and SBA-15 [14, 15]. In the first case, it
tion; (ii) preferably, a microporous structure, whereby a large has been showed that about 97.3% of RhB can be decom-
portion of the surface is inaccessible for large molecules posed (based on bleaching of RhB solution degree) when the
of organic substrates and diffusion difficulties slow down composition ­SnO2/porcine bone as a photocatalyst was used,
the photocatalytic transformations. There are several ways while only 51.5% of RhB can be degraded using pure ­SnO2
to overcome these shortcomings. Among them, introduc- nanoparticles. In the second case, the rate constant of RhB
tion of intrinsic (structural) and extrinsic (dopant) defects in photocatalytic degradation is 3.8 × 10−5 s−1; in the third case,
“white” oxides results in the band gap narrowing and mani- the rate constant achieves value of 1.5 × 10−4 s−1. Therefore,
festation of activity under visible irradiation [7]. The use of the possibility of pollutants photodegradation under visible
mechano- and sonochemical processes for these purposes irradiation using ­SnO2 embedded into structure of support,
as it was done for titania [8–10] and mixed oxides [11, 16],
* S. Khalameida was not studied.
svkhal@ukr.net At the same time, it is known that deposition of tin diox-
ide onto silica promotes to improvement of its catalytic per-
1
Institute for Sorption and Problems of Endoecology, NAS formance in acid-catalyzed and oxidation processes [17–21].
of Ukraine, Naumova Street 13, Kiev 03164, Ukraine
Several methods for deposition of S ­ nO2 on porous supports
2
L.V. Pysarzhevsky Institute of Physical Chemistry, NAS are known from the literature: sol–gel [22], polymeric [20],
of Ukraine, Nauky Ave. 31, Kiev 03028, Ukraine

13
Vol.:(0123456789)

2132 S. Khalameida et al.

template [21], and impregnation [17]. We used the simplest apparatus (F. Paulik, J. Paulik, L. Erdey) in the temperature
of them, namely impregnation of commercial silica gel and range of 20–1000 °C at a heating rate of 10° min−1. The ini-
subsequent precipitation or thermolysis of prepared precur- tial sample mass was about 200 mg, and the sensitivity was
sors. Silica being a dielectric was chosen as a support. 50 mg. Raman spectra were recorded using spectrograph
The aim of this work is the study of crystal and porous of Renishaw system (Ar laser, 514 nm). Strength (H0) and
structure, electronic characteristics of the supported ­SnO2/ acid sites concentration in ­SnO2/SiO2 catalysts were deter-
SiO2 compositions of different ­SnO2 content and influence mined using an indicator method and the back titration of
of these physicochemical properties on their photocatalytic n-butylamine by hydrochloric acid in the presence of bro-
activity under visible irradiation. mothymol blue, respectively, as well as using temperature-
programmed desorption (TPD) of ammonia [19]. The porous
structure of the initial silica gel and supported samples was
Experimental studied using nitrogen adsorption–desorption technique. The
isotherms were obtained using an automatic gas adsorption
Reagents and materials analyzer ASAP 2405 N (“Micromeritics Instrument Corp”)
after outgassing the catalysts at 150 °C for 2 h. The specific
Silica gel KSKG (China) was used as a support. In order to surface area S, mesopore volume Vme, and micropore vol-
increase the mesopore size, it was also subjected to hydro- ume Vmi were calculated from these isotherms using BET,
thermal treatment (HTT) in a vapor phase at 150 °C for 3 h BJH, and t-plot methods, respectively. The total pore volume
[23, 24]. Tin tetrachloride S
­ nCl4·5H2O was used as a source VΣ was determined by impregnation of the samples, which
of tin, aqueous ammonia solution—as a precipitant. were preliminarily dried at 150 °C, with liquid water (so-
called incipient wetness method [25]). Macropore volume
Preparation of supported ­SnO2/SiO2 compositions Vma was calculated as the difference between VΣ and sorp-
tion pore volume Vs. The latter one was determined from
Samples of ­SnO2/SiO2-supported compositions with differ- the isotherms at nitrogen relative pressure close to 1.0.
ent tin dioxide content, designated as XSn, where X = 1, 3, 5, The mesopore size dme was calculated from the curves of
7 and 10, (the number in the sample designation corresponds pore size distribution (PSD) plotted using the desorption
to ­SnO2 content in mass%), were synthesized by deposition branches of isotherms. Diffuse reflectance UV–Vis spec-
of tin dioxide on silica gel granules using incipient wetness tra of the powders were registered on Lambda 35 UV–Vis
impregnation. For this purpose, a fraction (0.5–2 mm) of spectrometer (PerkinElmer Instruments). The band gap was
silica gel was impregnated with an appropriate amount of determined using Planks formula.
­SnCl4·5H2O aqueous solution, held for 1 h, then treated
with 1 M N ­ H4OH solution to form an insoluble ­SnO2, then
dried at 110 °C and calcined in air at 450 and 550 °C (2 h). Photocatalytic testing
This is precipitation method (designation of samples − XSn-
prec). In another procedure, the stage of precipitation using Photocatalytic degradation was performed in a glass reac-
­NH4OH was absent. This is thermolysis method (designation tor under visible irradiation. LED Cool daylight lamp,
of samples − XSn-therm). For the purpose of activation, a Philips (100 W) possessing emission spectra exclusively
precipitated sample containing 5% S ­ nO2 was also milled in in the visible range with a broad maximum in the region
air for 0.5 h at 300 rpm (sample 5Sn-mill). For comparison, of 500–700  nm and a local maximum around 440  nm,
the supported sample was also prepared by direct dry milling was used as an irradiation source. The dye rhodamine
of mixture of 5% ­SnO2 and silica gel under the same condi- B (RhB) in the form of 1.5 × 10−5 mol ­L−1 solution was
tions (sample 5Sn-mill-mix). Milling was carried out using used as a pollutant [16]. The main absorption bands λmax
a planetary ball mill Pulverisette-7, premium line (Fritsch in the spectra of substrate is 553 nm. The catalyst dose
Gmbh) with a vessel of silicon nitride. Twenty-five balls was 1 g L−1 (80 mg of catalysts and 80 mL of solution).
from ­S3N4 with a 10 mm diameter (total ball mass − 40 g) Duration of dark adsorption to establish of equilibrium
were used as working bodies. was 60 min. The initial solution and solutions after dye
adsorption and degradation for 30–600 min were analyzed
Physicochemical measurements spectrophotometrically at λmax (Lambda 35, Perkin Elmer
Instruments) after centrifugation of the reaction mixture
The crystal structure of the supported samples was studied (10 min at 8000 rpm). The calculation of photodegrada-
by means of X-ray powder diffraction (XRD) using Philips tion rate Kd was based on the temporal changes of the dye
PW 1830 diffractometer with ­CuKα radiation. The curves concentration after reaching the adsorption equilibrium.
of DTA and TG were recorded using the Derivatograph-C The total organic carbon (TOC), which is a measure of dye

13
Physicochemical and photocatalytic properties of tin dioxide supported onto silica gel 2133

mineralization [26], for selected solutions, was determined well-consistent with those ones calculated by the formula
using a Shimadzu TOC analyzer (model 5050A). (2).
Precursors, prepared with a precipitation stage, contain
tin dioxide and ammonium chloride which are formed in the
pores of silica gel according to the equation:
Results and discussion
SnCl4 + 4NH3 + 2H2 O = SnO2 + 4NH4 Cl (3)
Thermogravimetric measurements In line with it, mass loss in the range of 200–450 °C for
these precursors is determined by the ammonium chloride
In order to determine the temperature of thermodestruc- sublimation [30] and the removal of hydroxyl groups both
tion of S
­ nCl 4·5H 2O resulting in tin dioxide formation, for silica gel and precipitated tin dioxide [1, 4, 5]. The
the DTA–TG curves for bulk salt as well as for compo- ­NH4Cl content in the products of the reaction (3) is about
sitions precursors prepared on its basis were registered. 58.6% w/w and mass loss in the range of 200–450 °C for
Obviously, ­SnCl4·5H2O is decomposed in air according precipitated ­SnO2 is 4.86% [4]. Therefore, total mass loss
to Eq. (1): in the range of 200–450 °C for the precursors prepared via
precipitation can be calculated as follows:
SnCl4 ⋅ 5H2 O = SnO2 + 4HCl + 3H2 O (1)
Δm200−450_prec = X ⋅ 58.6∕100 + X ⋅ 4.86∕100 + (1 − X) ⋅ 0.72∕100
Theoretical mass loss for this process is 54.1% w/w.
(4)
Experimental value of mass loss calculated from DTA–TG
curves of bulk S ­ nCl 4·5H 2O is 55.5% w/w. At the same The experimental data presented in column 7 of Table 1
time, this process takes place in a wide temperature are in a good agreement with those ones calculated by
range—from 100 to 450 °C (Fig. 1a). It can be divided into the formulas (2) and (4).
two stages: (i) sharp stage at 100–200 °C with an intensive Based on the results of thermogravimetric measure-
endoeffect at about 150 °C when mass loss is 47.9% w/w ments, the first calcination temperature of precursors was
(crystallization water) and (ii) slow stage at 200–450 °C determined as 450 °C. The second calcination temperature
when mass loss is 7.6% w/w. It should be noted that simi- of 550 °C was chosen since content of surface OH groups
lar curves were obtained for hydrous copper, cobalt, and was reduced by 11% compared with the samples calcined at
manganese chlorides [27]. 450 °C for silica gel as shown in [21] and by 8% for tin diox-
On the other hand, there is an endoeffect at 20–200 °C ide xerogel [4] according to the data of DTA–TG. Therefore,
with a maximum at 115  °C, which is accompanied by use of the samples calcined at these temperatures makes
release of water from the pores, for the initial porous silica it possible to establish the influence of the content of OH
and supported samples, namely their precursors (Fig. 1b, c, groups on the photocatalytic activity. It should also be noted
respectively). These endoeffects are characteristic for porous that the indicated values of temperature (450 and 550 °C)
materials including supported compositions based on porous correspond to optimal conditions for regeneration of oxide
carriers [11–13, 16, 28, 29]. Therefore, the first stage of catalysts using oxygen and air, respectively [31].
­SnCl4·5H2O decomposition for supported precursors is over-
lapped with removal of this physically bound water from Crystal structure
the pores of silica gel. At the same time, removal of surface
OH groups additionally occurs on the second stage which Crystal structure of tin dioxide in composition is formed
is the characteristic for silica [23]. This value is 0.72% w/w during calcinations of precursors. XRD patterns for some
(Table 1). supported samples as well as for pure S ­ nO2 calcined at
Therefore, mass loss in the range of 200–450 °C for 550 °C are depicted in Fig. 2. As can be seen, positions
precursors is determined by the tin chloride decomposi- of the diffraction peaks corresponding to the planes (110),
tion process and the removal of hydroxyl groups. The first (101), (200), and (211) are attributed to tetragonal modifica-
component depends on the precursor composition, i.e., tion of cassiterite/rutile (JCPDS N 41-1445). The intensity
content of S­ nCl 4, and the second one is approximately of all the peaks is predictably increased with an increase in
constant—about 0.72% w/w: the content of tin dioxide. The reflexes also differ in width,
Δm200−450_therm = X ⋅ 7.6∕100 + (1 − X) ⋅ 0.72∕100 (%w/w), which is a measure of crystallinity (crystallite size Dhkl),
(2) and I110/I101 ratio. Since reflex with maximal intensity (110)
where X—SnO2 content in composition. is overlapped with halo characteristic for amorphous silica,
It can be seen that Δm200–450_therm values, determined calculation of crystallites size was performed using reflex
from the experimental data (Table  1, column 6), are attributed to the plane (101) (D101). These parameters are
determined by ­SnO2 content in composition and calcinations

13

2134 S. Khalameida et al.

Fig. 1  a DTA, TG, DTG curves DTG/mg min–1


for ­SnCl4·5H2O. b DTA, TG,
(a) Mass loss/% DTA/µV mg–1
DTG curves for the initial silica
0.6 0.2
gel. c DTA, TG, DTG curves TG
for a precursor containing 10% 0
­SnCl4 DTG 0.5 0.0

12 0.4
– 0.2

0.3
24 – 0.4
DTA
0.2
36 – 0.6

0.1
– 0.8
48
0.0
– 1.0
60 – 0.1
– 1.2
200 400 600 800
Temperature/°C

DTG/mg min–1
(b) Mass loss/%
DTA/µV mg–1 0.45
0
TG 0.40
0.0
1
DTG 0.35

2 0.30

DTA 0.25
3
– 0.1

4 0.20

0.15
5

0.10
6 – 0.2
0.05
200 400 600 800
Temperature/°C

DTG/mg min–1
(c) Mass loss/% DTA/µV mg–1
0.50
TG
0
0.45
1
0.40
DTA
0.0
2 0.35

3 DTG 0.30

4 0.25

0.20 – 0.1
5

0.15
6
0.10
7
0.05 – 0.2
200 400 600 800
Temperature/°C

13
Physicochemical and photocatalytic properties of tin dioxide supported onto silica gel 2135

Table 1  Results of N Synthesis conditions Temperature of effect Δm20–200/% w/w Δm200–450/% Δm450–550/%


thermogravimetric from DTG curve/°C w/w w/w
measurements for some ­SnO2/
SiO2 samples 1 Initial silica gel 113 3.71 0.72 0.19
2 3Sn without precipitation 110 3.88 0.93 0.20
3 5Sn without precipitation 115 3.60 1.13 0.21
4 5Sn with precipitation 118 4.11 3.94 0.26
5 7Sn without precipitation 113 3.69 1.31 0.22
6 10Sn without precipitation 118 3.53 1.48 0.24
7 10Sn with precipitation 121 4.55 6.88 0.29
8 Precipitated ­SnO2 128 13.01 4.86 0.48

1200 Table 2  Some parameters of crystal structure


900 Sample Method of prepara- Temperature/°C D101/nm I110/I101
a
600 tion
300 3Sn Precipitation 450 5.0 1.69
0 550 5.8 1.55
3300 5Sn Precipitation 450 6.0 1.88
b 550 7.1 1.70
2200
Thermolysis 450 7.3 1.58
1100
550 10.0 1.50
0
7Sn Precipitation 450 7.5 1.55
3300 550 8.8 1.45
Intensity/a.u.

c 10Sn Precipitation 450 10.4 1.52


2200

1100
550 11.0 1.40
Thermolysis 550 11.2 1.34
0
100Sn Precipitation 450 12.0 1.05
3300
d 550 15.6 1.11
2200

1100

0
110 101
A feature of the supported samples also is an increase
17700
in I110/I101 value (Table 2, column 5) compared to that for
11800 e pure ­SnO2 prepared under the same conditions as well as
5900 described in the literature [1, 6, 33]. It can also be seen that
0 this value is larger for the samples prepared via precipitation
10 15 20 25 30 35 40 45 50 55 60 and decreases when calcinations temperature is elevated.
2θ /° This parameter characterizes the surface orientation of
defined planes which can be very important for catalytic
Fig. 2  XRD patterns for the supported samples after calcinations at processes [34, 35] including photocatalytic transformation
450 °C: 3Sn_prec (a), 5Sn_therm (b), 5Sn_prec (c), 10 Sn_prec (d), [36]. Particularly, the  ratio of different crystallographic
bulk precipitated ­SnO2 (e) planes (surface content) and, as a result, catalytic proper-
ties, could be changed by various types of treatment for
temperature which is also quite expected (Table 2). Thereby, vanadium–molybdenum oxides [35, 37]. These are so-called
­D101 value is increased as a result of enrichment of the com- structural-sensitive reactions [1, 36].
position with tin dioxide and an elevation in the calcinations Raman spectra confirm low crystallinity for deposited
temperature. On the other hand, calculated crystallite size ­SnO2 phase which can be seen from the comparison of the
for the supported samples is 2–3 times smaller than for bulk spectra registered for bulk and supported tin dioxide (Fig. 3)
­SnO2 calcined at 450–550 °C. It indicates well-dispersion As can be seen (Fig. 3, spectra b, c), there are low-inten-
of the deposited phase on the support surface as it was pre- sity bands centered at 795 and 590 cm−1 which attributed
viously observed for supported titania [8–10] as well as for to ­SnO2 nanoparticles with a size less than 10 nm [17, 38,
mixed oxide compositions [11, 32]. 39]. Besides, broad bands at 300–550 cm−1, which can be a

13

2136 S. Khalameida et al.

superposition of several bands characteristic for tin dioxide


4000 a [39], are presented in Fig. 3 (spectra b and c). Similar spectra
3000 were also obtained for samples containing 5–10% of ­SnO2
incorporated into structure of dealuminated zeolite ß [18]. It
2000 should be noted that there are no bands for amorphous silica
in this range [40].
1000

0
Porous structure
750
b
Parameters of porous structure for initial silica gel and
Intensity/a.u.

500 supported samples were calculated from nitrogen adsorp-


tion–desorption isotherms, examples of which are shown in
250 Fig. 4. All of them belong to type IV according to IUPAC
classification which is characteristic for mesoporous materi-
0
als. The curves of pore size distribution PSD (inset of Fig. 4)
1000 also indicate it.
c Thus, initial silica gel contains mesopores with a size
750 in the range of 2–14  nm and maximum on PSD curve
about 7.8 nm but does not contain micro- and macropores
500 (Table 3). Deposition of tin dioxide on siliceous surface
predictably leads to some decrease in the specific surface
250 area, total pore volume and mesopore volume but their size
remains almost the same (inset of Fig. 4, Table 3, columns
0 7). Silica gel subjected to HTT has a smaller specific sur-
0 500 1000 1500 2000
face area but larger mesopores, namely 9.8 nm. Therefore,
Raman shift/cm–1
the supported sample based on the hydrothermally treated
silica gel also has a smaller specific surface area and larger
Fig. 3  Raman spectra for bulk tin dioxide (a) and supported samples
containing 5 (b) and 10% (c) of S
­ nO2 mesopores. Maximum reduction in the specific surface area

Fig. 4  Nitrogen adsorption–
0.035
desorption isotherms and PSD Initial silica gel
curves for the initial silica gel
700 0.030
and some supported samples 1Sn-prec
0.025 5Sn-prec
600
0.020 5Sn-mill
dv/dr

0.015
500

0.010

400
a/cc g–1

0.005

0.000
300
0 5 10 15 20
r /nm

200

100

0
0.0 0.2 0.4 0.6 0.8 1.0
P/P0

13
Physicochemical and photocatalytic properties of tin dioxide supported onto silica gel 2137

Table 3  Parameters of porous Sample Method of preparation Temperature/°C S/m2 g−1 VΣ/cm3 g−1 Vme/cm3 g−1 dme/nm
structure of supported samples
SiO2 – – 371 0.97 0.98 7.8
1Sn Precipitation 450 365 0.92 0.92 7.8
Thermolysis 450 367 0.93 0.92 7.8
3Sn Thermolysis 450 362 0.90 0.90 7.8
550 359 0.91 0.90 7.8
5Sn Thermolysis 450 357 0.93 0.92 7.9
550 355 0.94 0.93 7.9
Precipitation 450 360 0.91 0.91 7.8
Precipitation 550 354 0.90 0.89 7.8
5Sn-HTT Precipitation 450 320 0.90 0.89 9.5
5Sn-mill Precipitation – 198 0.50 0.48 7.9
5Sn-mill-mix Precipitation – 301 0.74 0.73 7.8
7Sn Thermolysis 450 350 0.90 0.89 8.1
550 345 0.89 0.89 8.0
10Sn Thermolysis 450 344 0.88 0.87 8.1
550 340 0.87 0.87 8.1

and pore volume is observed for the milled sample 5Sn- UV–Vis spectra
mill (Fig. 4, Table 3, column 4) which is consistent with the
results obtained for a milled silica gel of different porous Diffuse reflectance UV–Vis spectra for some supported
structure [23, 41]. samples with different ­SnO2 content are presented in Fig. 5.
One can see that they differ in the position of the absorp-
tion edge λ and the magnitude of absorption in the visible
region. It is important that significant redshift of absorption

Fig. 5  UV–Vis spectra for some


supported samples 1.0 1Sn-prec
10Sn-prec
0.9 5Sn-prec
5Sn-therm
0.8

0.7
0.5
K – M index

0.6

0.5 0.4
K – M index

0.4
0.3

0.3

0.2
0.2
350 400 450
Wavenumber/nm
0.1
200 300 400 500

Wavenumber/nm

13

2138 S. Khalameida et al.

Table 4  Electronic and photocatalytic properties of the supported samples


N SnO2 content/% Preparation conditions λ/nm Eg/eV a/% Kd 105/s−1 Bleaching
degree/%

1 1Sn Precipitation, 450 °C 392 3.16 40 5.6 84


2 3Sn Precipitation, 450 °C 399 3.11 45 7.2 88
3 5Sn Precipitation, 450 °C 407 3.05 56 9.1 92
4 5Sn Precipitation, 550 °C 388 3.20 40 7.0 84
5 5Sn Thermolysis, 450 °C 375 3.31 44 6.1 80
6 5Sn Thermolysis, 550 °C 365 3.40 37 5.1 76
7 5Sn Precipitation, 450 °C, ­SiO2 treated via HTT *n.d – 33 3.9 63
8 5Sn Precipitation, 450 °C MChT *n.d – 32 3.1 68
9 5Sn Precipitation, 450 °C + SiO2 MChT *n.d – 37 0.1 11
10 7Sn Precipitation, 450 °C 367 3.38 41 5.8 82
11 10Sn Precipitation, 450 °C 364 3.41 25 3.1 59
12 10Sn Thermolysis, 450 °C 356 3.48 36 2.0 44
13 100Sn Precipitation, 450 °C 341 3.64 30 – –

*n.d not done

edge is observed for the supported samples compared with A feature of the spectrum for 5Sn-prec-450 is the pres-
bulk ­SnO2. ence of another absorption edge in the long-wavelength
Thus, λ is 364–341 nm for bulk tin dioxide [4, 5], while region—about 540 nm (selected fragment in Fig. 5). As a
its value is shifted toward 407–356 nm for the precipi- result, the occurrence of an additional level in the forbidden
tated samples containing 1–10% ­SnO2 (Table 4). It cor- zone of ­SnO2 is possible. Similarly, a new metastable surface
responds to narrowing the band gap Eg from 3.64 eV for phase exhibits with much reduced band gap from that one of
bulk ­SnO2 [4, 5] to 3.05–3.48 eV for the supported sam- bulk ­TiO2 (about 2 eV) on the rutile ­TiO2 (011) surface [36].
ples (Table 4, columns 4, 5). Such an effect was earlier
described for silica supported T ­ iO2 photocatalysts [8, 9] Photocatalytic tests
and for ­SnO2–TiO2 solid solution embedded into SBA-15
structure [42]. But it was not observed for deposited S ­ nO2, It is well-known that the photocatalytic properties of semi-
as a rule [15, 21, 42]. Similar spectrum is given only in the conductors depend on their electronic characteristics, crystal
paper [18] for sample containing 10% S ­ nO2 incorporated and porous structure, specific surface area and degree of
into the structure of dealuminated ß zeolite. However, the hydroxylation, and therefore, adsorption capacity toward
authors do not discuss it. Moreover, co-precipitated (but the substrate [7, 26]. It is also natural that the first stage of
not supported) ­SnO2–SiO2 samples prepared by a sol–gel
method and additionally calcined at 800 °C showed red-
shift in UV–Vis spectra with an increase in ­SnO2 content 1.0
from 1 to 10% [43]. 0.30

It is noteworthy that observed redshift of absorption 0.8


0.25
edge for the precipitated samples is significantly larger
than for the samples prepared via thermolysis (Fig. 5,
TCD signal/a.u.

0.20
0.6
Table 4). An observed redshift effect can be associated
dVHCl/dT

with different surface orientations of deposited crystallites 0.15

compared to crystallites in bulk S ­ nO2 (Table 2, increase in 0.10


0.4

I110/I101 ratio from 1.05–1.10 to 1.34–1.88). It is in a good


agreement with the results described for ­TiO 2 in [30]; 0.05 0.2
UV–Vis spectroscopy measurements of anatase crystals
0.00
with different surface orientations have revealed differ-
0.0
ences in the band gap indicating the upward shift in the 0 100 200 300 400 500 600 700
conduction band minimum for anatase (101) and (100) Temperature/°C
compared to (001) surfaces.
Fig. 6  NH3-TPD profile for the sample 5Sn_prec-450

13
Physicochemical and photocatalytic properties of tin dioxide supported onto silica gel 2139

catalysis is the adsorption of reagents on the catalyst surface. (a) 0.3 0


Based on the fact that the tested dye has a basic nature, it was 1800 s
important to evaluate the acidic properties of the photocata- 3600 s
7200 s
lyst surface. Using the indicator method, it was established 10,800 s
0.2
that the supported samples have concentration of acid sites 18,000 s
within 1.58–1.68 mmol g−1 or 4.3–4.6 μmol m−2. The value 27,000 s
36,000 s

D
of Gammet function H0 is about − 3, i.e., acid sites are weak. RhB initial
At the same time, ammonia TPD profile obtained for the 0.1

sample 5Sn_prec-450 (Fig. 6) shows the value of acid sites


concentration equal to 1.51 mmol g−1 which is close to the
above indicated range. Temperature of the maximum is at 0.0
172 °C which also corresponds to weak acid sites. It should
be noted that similar results were obtained for ­SnO2–SiO2 400 500 600

xerogels containing 3–9% w/w of tin dioxide [44], while Wavenumber/nm


content of acid sites for bulk ­SiO2 and ­SnO2 is 1.10 and
(b) 0.30
0.38 mmol g−1 or 3.0 and 6.5 μmol m−2, respectively. As 0
can be seen, the supported samples have close acidity val- 0.25 1800 s
ues, but higher than the values based on the additivity rule. 3600 s
7200 s
Therefore, the difference in RhB adsorption during dark 0.20
10,800 s
stage a (%), determined as the ratio of optical density of 18,000 s
0.15
RhB solution before and after dark stage, depends primarily 27,000 s
D 36,000 s
on the value of the specific surface area and, possibly, on the 0.10
surface content of (110) plane (I110/I101 ratio).
At the same time, band gap (namely < 3.26  eV) and 0.05
absorption value in the region with wavelength > 380 nm for
the photocatalyst determine its activity under visible irradia- 0.00

tion [7]. As shown in Table 4, some of the prepared samples


200 300 400 500 600 700
(namely NN 1-4) meet these conditions. Consequently, the Wavenumber/nm
appearance of an electron–hole pair on the catalyst surface
due to the action of photons of visible light is possible in this Fig. 7  a Spectra of RhB depending on degradation time in the pres-
case. It results in initiation of a conventional photocatalytic ence of 5Sn_prec_450. b Spectra of RhB depending on degradation
process. As a result, these samples showed high activity in time in the presence of 3Sn_prec_450
the process of RhB degradation. The latter is evidenced by
the character of the changes in the RhB spectrum during
its degradation in the presence of the samples of different shift of this band toward 498–500 nm for 180 min which is
activity (Fig. 7a, b). The similar spectra were also obtained accompanied by complete de-ethylation of RhB to Rh110;
for other photocatalysts. It should be noted that bulk pre- (iii) decrease in intensity of the band at 498–500 nm which
cipitated ­SnO2 prepared under the same conditions does not is associated with degradation of Rh110 with cleavage of the
show noticeable activity when it is exposed to visible light chromophore rings (for duration > 180 min).
[4] which is associated with a high band gap value (3.41 eV) The photocatalytic tests showed that the degradation
and minimal absorption in visible region for this sample. of RhB under visible irradiation (as dye bleaching) using
It is well-known that RhB photodegradation in the pres- the supported samples proceeds according to first-order
ence of tin dioxide can occur by two pathways [4–6]: as de- kinetic equation (Fig.  8) with a correlation coefficient
ethylation process in a stepwise manner (with the formation R2 = 0.93–0.99, which is agreed with literature data [4–6,
of three intermediates from RhB to Rh110) or as a direct 16]. The photodegradation rate constants Kd were calculated
cleavage of the chromophore rings. A gradual blueshift of from the slopes of the plots ln(D/D0) − τ (where D and D0
the band at 553 nm on the spectrum of the RhB solution is are values of optical density of RhB solution after time τ
observed in the first case and the reduction in its intensity and dark adsorption, respectively). One can see (Table 4)
without shift takes place in the second case. As shown in that photocatalytic activity, the measure of which is the Kd
Fig. 7a, b, mixed mechanism is realized in this case as it magnitude, for the precipitated samples is increased with
was described earlier for doped and milled ­SnO2 [4, 5]: (i) an increase in S­ nO2 content up to 5% and decreased with a
decrease of the intensity band at 553 nm occurs for 60 min further increase in its content up to 10%. In general, it corre-
which corresponds to direct degradation of RhB [16]; (ii) lates with the electronic characteristics of the photocatalysts:

13

2140 S. Khalameida et al.

0.5 Particularly, it is difficult to associate an increase in crys-


0.0 tallite size D101, which is observed with increasing S ­ nO2
content (Tables 2, 4), with activity. Here, it is important to
– 0.5
have an actual increase in the content of the active phase,
– 1.0 as established for 5.8–17.7% of S ­ nO2 embedded into SBA-
15 under UV irradiation [15]. At the same time, surface
In/D/D0

– 1.5
exposition of certain crystallographic planes can affect the
– 2.0
SnO2 bulk dye adsorption and through it on photocatalytic activity. On
– 2.5 1Sn/SiO2 the contrary, the dependence of the activity on the specific
5Sn/SiO2
– 3.0
surface area was precisely established earlier [1, 7]. At the
10Sn/SiO2
same time, the tested samples have close S values (Table 3).
– 3.5
In order to reduce the specific surface area of the support,
– 4.0 and therefore the supported sample, silica gel was subjected
0 5000 10000 15,000 20,000 25,000 30,000 35,000 40,000
to HTT. Photocatalyst prepared on the basis of a hydrother-
τ /s
mally treated support showed significantly lower activity
(Table 4, sample 7).
Fig. 8  Kinetic curves of RhB degradation for some tested samples Finally, an attempt to activate the most effective cata-
lyst (sample 3) by milling, as was done for bulk S ­ nO2 [5],
the samples containing 1–5% ­SnO2 have Eg = 3.05–3.20 eV, was made. As shown in Table 4 (sample 8), positive effect
i.e., less than 3.26 eV and absorption in visible region for was not achieved. Furthermore, the samples prepared via
them is 25–40%, while these parameters are 3.35–3.41 eV milling of tin dioxide and silica gel mixture (without pre-
and 15–20%, respectively, for samples containing 7–10% liminary deposition of ­SnO2 on silica gel) was practically
­SnO2 (Table 4). Therefore, generation of an electron–hole inactive, although the adsorption of RhB was quite large
pair is possible in the first case. Moreover, the number of (sample 9). It can be explained as follows. Under milling,
active sites should be increased with an increase in ­SnO2 ­SnO2 (semiconductor) is partially (sample 8) or completely
content. In the second case, the samples also exhibit suffi- (sample 9) coated by a silica layer (dielectric). As a result,
cient photocatalytic activity. The latter can be explained as the active sites of S
­ nO2 are blocked. Taking into account a
follows. The spectrum of RhB absorption is overlapped in a large excess of silica in relation to S­ nO2, as well as the dif-
wide region with the emission spectrum of a lamp—illumi- ference in the hardness of the components (Mohs hardness
nation source; the maxima are located at 553 and 565 nm, of amorphous silica and S ­ nO2 is 5 and 6.5, respectively)
respectively. Therefore, the absorption of visible light and [46, 47], “smearing” of amorphous silica on S ­ nO2 nano-
the direct excitation of RhB molecules are very likely. Sub- crystallites is possible. Similar results were described when
sequent injection of an electron from the excited RhB mol- “core–shell” particles were formed during dry milling of
ecules into the conduction band of a photocatalyst occurs. It cerium and molybdenum oxides mixtures [48].
is so-called photosensitization contributing the initiation of As shown in Table 4 (column 8) bleaching degree of RhB
photocatalytic process [26, 45]. Obviously, this mechanism under visible irradiation for the most active sample reaches
determines the photocatalytic activity of the samples pos- 88–92% which is comparable with ones obtained in [14]
sessing Eg > 3.26 eV, although it can also contribute to the for the same dye using the ­SnO2/porcine bone photocatalyst
activity of the samples with Eg < 3.26 eV. but under UV irradiation. At the same time, the mineraliza-
For the most active sample, namely containing 5% S ­ nO2, tion degree of a dye is even more important indicator for
the influence of preparation conditions on the photocatalytic evaluating of the photocatalyst efficiency [26]. This param-
activity was studied. One can see that the samples prepared eter, calculated as the degree of total organic carbon (TOC)
via thermolysis have lower activity, and it is directly related reduction, is 40 and 75% for sample containing 5% S ­ nO2
to the larger Eg value and lower absorption for visible irra- calcined at 450 °C and prepared via thermolysis and pre-
diation (Table 4). An increase in calcinations temperature cipitation, respectively.
for the samples prepared via precipitation and thermolysis
from 450 to 550 °C results in a slight decrease in activity. It
can be explained by a decrease in concentration of the sur- Conclusions
face OH groups which influence on photocatalytic processes.
As mentioned above, their content according to TG data is Using thermogravimetric analysis, it has been established
reduced by approximately 10% in this temperature range. that hydrous tin tetrachloride is transformed into tin dioxide
It is also important to evaluate the effect of other phys- in silica gel pores at 450 °C. The choice of the temperature
icochemical characteristics on photocatalytic properties. of the precursor transformation into the oxide composition

13
Physicochemical and photocatalytic properties of tin dioxide supported onto silica gel 2141

is based on these results. Also, it has been estimated the physical–chemical, and catalytic properties of mixed compositions
degree of surface dehydroxylation within the range of Ag/H3PMo12O40/SiO2. J Therm Anal Calorim. 2011;107:453–61.
13. Khalameida S, Sydorchuk V, Skubiszewska-Zięba J, Leboda R,
450–550 °C. Deposition of 1–10% tin dioxide onto silica Zazhigalov V. Glass Phys Chem. 2014;40:8–16.
gel contributes to its dispersion on the surface; crystallite 14. Wu Y, Wang H, Cao M, Zhang Y, Cao F, Zheng X, Hu J, Dong
size of ­SnO2 is decreased from 12 nm for a bulk sample to J, Xiao Z. Animal bone supported S ­ nO2 as recyclable photocata-
5–10 nm (depending on ­SnO2 content) for the supported lyst for degradation of rhodamine B dye. J Nanosci Nanotechnol.
2015;15:6495–502.
samples. Besides, surface re-orientation of the crystallites 15. Srinivasan NR, Majumdar P, Eswar NKR, Bandyopadhyaya R.
occurs: content of (110) plane, expressed as I110/I101 ratio, Photocatalysis by morphologically tailored mesoporous silica
is increased by 45–80%. Moreover, narrowing the band gap (SBA-15) embedded with ­SnO2 nanoparticles: experiments and
from 3.64 eV for bulk ­SnO2 to 3.05–3.48 eV for the samples model. Appl Catal A. 2015;498:107–16.
16. Dippong T, Levei E, Cadar O, Goga F, Toloman D, Borodi G.
containing 1–10% ­SnO2 is observed. As a result, the sup- Thermal behavior of Ni, Co and Fe succinates embedded in silica
ported samples, in contrast to the bulk S­ nO2, exhibit photo- matrix. J Therm Anal Calorim. 2019;136:1587–95.
catalytic activity under visible irradiation. The precipitated 17. Ferrini P, Dijkmans J, De Clercq R, Van de Vyver S, Dus-
samples are more effective compared with the samples pre- selier M, Jacobs PA, Sels BF. Lewis acid catalysis on single
site Sn centers incorporated into silica hosts. Coord Chem Rev.
pared by thermolysis. An increase in calcinations tempera- 2017;343:220–55.
ture decreases the photocatalytic activity due to a reduction 18. Hammond C, Conrad S, Hermans I. Simple and scalable prepa-
in the specific surface area and surface dehydroxylation. ration of highly active lewis acidic Sn-b. Angew Chem Int Ed.
The precipitated sample with 5% of tin dioxide calcined at 2012;51:11736–9.
19. Varvarin A, Levytska S, Brei V. Conversion of ethyllactate into
450 °C, which has I110/I101 = 1.88 and band gap 3.05 eV, lactide over acid ­SnO2/SiO2 catalyst. Kataliz i neftechimia (Russ).
possesses maximal activity. 2018;27:19–23.
20. Carreño NLV, Nunes MR, Raubach CW, Granada RL, Krolow
MZ, Orlandi MO, Fajardo HV, Probst LFD. ­SnO2 nanoparticles
functionalized in amorphous silica and glass. Powder Technol.
References 2009;195:91–5.
21. Skoda D, Styskalik A, Moravec Z, Bezdicka P, Bursik J, Mutine
1. Batzill M, Diebold U. The surface and materials science of tin PH, Pinkas J. Mesoporous ­SnO2–SiO2 and Sn–silica–carbon nano-
oxide. Progress Surf Sci. 2005;79:47–154. composites by novel non-hydrolytic templated sol–gel synthesis.
2. Samsonenko M, Zakutevskyy O, Khalameida S, Charmas B, RSC Adv. 2016;6:68739–47.
Skubiszewska-Ziȩba J. Influence of mechanochemical and micro- 22. Cai J, Li Z, Yao S, Meng H, Shen PK, Wei Z. Close-packed ­SnO2
wave modification on ion-exchange properties of tin dioxide with nanocrystals anchored on amorphous silica as a stable anode mate-
respect to uranyl ions. Adsorption. 2019;25:451–7. rial for lithium–ion battery. Electrochim Acta. 2012;74:182–8.
3. Wu J, Zeng D, Tian S, Xu K, Li D, Xie C. Competitive influence 23. Skubiszewska-Zięba J, Khalameida S, Sydorchuk V. Comparison
of surface area and mesopore size on gas-sensing properties of of surface properties of silica xero- and hydrogels hydrothermally
­SnO2 hollow fibers. Mater Sci. 2015;50:7725–34. modified using mechanochemical, microwave and classical meth-
4. Khalameida S, Samsonenko M, Sydorchuk V, Starchevskyy V, ods. Colloids Surf A. 2017;504:139–53.
Zakutevskyy O, Khyzhun O. Theor Exp Chem. 2017;53:40–5. 24. Leboda R, Mendyk E, Tertykh VA. Effect of the hydrothermal
5. Khalameida S, Samsonenko M, Skubiszewska-Ziȩba J, Zakutevs- treatment method in an autoclave on the silica gel porous struc-
kyy O. Dyes catalytic degradation using modified tin(IV) oxide ture. Mater Chem Phys. 1995;42:7–11.
and hydroxide powders. Adsorpt Sci Technol. 2017;35:853–65. 25. Leofanti G, Padovan M, Tozzola G, Venturelli B. Surface area and
6. Sangami G, Dharmaraj N. UV–visible spectroscopic estimation pore texture of catalysts. Catal Today. 1998;41:207–19.
of photodegradation of rhodamine-B dye using tin(IV) oxide 26. Rauf MA, Ashraf SS. Fundamental principles and application
nanoparticles. Spectrochim Acta Part A Mol Biomol Spectrosc. of heterogeneous photocatalytic degradation of dyes in solution.
2012;97:847–52. Chem Eng J. 2009;151:10–8.
7. Hoffmann MR, Martin ST, Choi W, Bahneman DW. Environ- 27. Yi X, Hu J, Sun M, Man X, Zhang Y, Liu S. Thermal stability
mental applications of semiconductor photocatalysis. Chem Rev. and decomposition behaviors of some hydrous transition metal
1995;95:69–96. chlorides. J Therm Anal Calorim. 2019. https​://doi.org/10.1007/
8. Xu Y, Zheng W, Liu W. Enhanced photocatalytic activity of sup- s1097​3-019-08307​-4.
ported ­TiO2: dispersing effect of ­SiO2. J Photochem Photobiol A 28. Catauro M, Naviglio D, Risoluti R, Ciprioti SV. Sol–gel synthesis
Chem. 1999;122:57–60. and thermal behavior of bioactive ferrous citrate–silica hybrid
9. Ding Z, Hu X, Lu GQ, Yue P-L, Greenfield PF. Novel silica gel materials. J Therm Anal Calorim. 2018;133:1085–92.
supported ­TiO2 photocatalyst synthesized by CVD method. Lang- 29. Dippong T, Levei EA, Cadar O, Goga F, Borodi G, Barbu-Tudoran
muir. 2000;16:6216–22. L. Thermal behavior of ­C oxFe 3−xO 4/SiO 2 nanocomposites
10. Van Grieken R, Aguado J, López-Muñoz MJ, Marugán J. Syn- obtained by a modified sol–gel method. J Therm Anal Calorim.
thesis of size-controlled silica-supported ­TiO2 photocatalysts. J 2017;128:39–52.
Photochem Photobiol A Chem. 2002;148:315–22. 30. Olszak-Humienik M. On the thermal stability of some ammonium
11. Sidorchuk V, Tertykh V, Klimenko V, Ragulya A. Formation and salts. Thermochim Acta. 2001;378:107–12.
some properties of barium titanate embedded into porous matri- 31. Argyle MD, Bartholomew CH. Heterogeneous catalyst deactiva-
ces. J Therm Anal Calorim. 2010;101:729–35. tion and regeneration: a review. Catalysts. 2015;5:145–269.
12. Trach Y, Sydorchuk V, Makota O, Khalameida S, Leb- 32. Ştefănescu M, Muntean C, Berei E, Ştefănescu O. Prepara-
oda R, Skubiszewska-Zięba J, Zazhigalov V. Synthesis, tion and characterization of ­CuCr2O4/SiO2 and ­Cu2Cr2O4/SiO2

13

2142 S. Khalameida et al.

nanocomposites obtained from carboxylate complex combina- activation in various media on structure of porous and non-porous
tions. J Therm Anal Calorim. 2019. https:​ //doi.org/10.1007/s1097​ silicas. Appl Surf Sci. 2010;257:446–50.
3-019-08479​-z. 42. Srinivasan NR, Bandyopadhyaya R. ­SnxTi1−xO2 solid-solution-
33. Vladut CM, Mihaiu S, Szilágyi IM, Kovács TN, Atkinson I, nanoparticle embedded mesoporous silica (SBA-15) hybrid as an
Mocioiu OC, Petrescu S, Zaharescu M. Thermal investigations engineered photocatalyst with enhanced activity. Faraday Discuss.
of the Sn–Zn–O gels obtained by sol–gel method. J Therm Anal 2016;186:353–70.
Calorim. 2019;136:461–70. 43. del Castillo J, Yanes AC, Méndez-Ramos J, Rodríguez VD. Lumi-
34. Si R, Flytzani-Stephanopoulos M. Shape and crystal-plane effects nescence of nanostructured S ­ nO2–SiO2 glass-ceramics prepared
of nanoscale ceria on the activity of Au-CeO2 catalysts for the by sol–gel method. J Nanosci Nanotechnol. 2008;8:2143–6.
water–gas shift reaction. Angew Chem Int Ed. 2008;47:2884–7. 44. van Grieken R, Martos C, Sánchez-Sánchez M, Serrano DP,
35. Skwarek E, Khalameida S, Janusz W, Sydorchuk V, Konovalova Melero JA, Iglesias J, Cubero JG. Synthesis of Sn–silicalite from
N, Zazhigalov V, Skubiszewska-Zięba J, Leboda R. Influence of hydrothermal conversion of S ­ iO2–SnO2 xerogels. Microporous
mechanochemical activation on structure and some properties of Mesoporous Mater. 2009;119:176–85.
mixed vanadium–molybdenum oxides. J Therm Anal Calorim. 45. Wu T, Liu G, Zhao J, Hidaka H, Serpone N. Photoassisted deg-
2011;106:881–94. radation of dye pollutants. V. Self-photosensitized oxidative
36. Batzill M. Fundamental aspects of surface engineering of transformation of Rhodamine B under visible light irradiation in
transition metal oxide photocatalysts. Energy Environ Sci. aqueous ­TiO2 dispersions. J Phys Chem B. 1998;102:5845–51.
2011;4:3275–86. 46. Davraz M, Gunduz L. Engineering properties of amorphous silica
37. Khalameida S, Sydorchuk V, Leboda R, Skubiszewska-Zięba J, as a new natural pozzolan for use in concrete. Cem Concr Res.
Zazhigalov V. Physical-chemical transformations in the system 2005;35:1251–61.
­V2O5/(NH4)2Mo2O7 under hydrothermal conditions. Cent Eur J 47. Wood GC, Hodgkiess T. The hardness of oxides at ambient tem-
Chem. 2014;12:140–52. peratures. Mater Corros. 1972;23:766–73.
38. Dieguez A, Romano-Rodrıguez A, Vila A, Morante JR. The com- 48. Zazhigalov V, Sachuk O, Diyuk O, Bacherikova I, Posudievsky
plete Raman spectrum of nanometric S ­ nO2 particles. J Appl Phys. O, Shcherban N. Mechanochemical synthesis of nanosized com-
2001;90:1550–7. pounds in ­CeO2–MoO3 system. In: Proceedings of 2018 IEEE 8th
39. Yu KN, Xiong Y, Liu Y, Xiong C. Microstructural change of international conference “nanomaterials: applications & proper-
nano-SnO2 grain assemblages with the annealing temperature. ties”. 01SPN35-1-01SPN35-6.
Phys Rev B Condens Matter. 1997;55:2666–71.
40. Volovšek V, Furić K, Bistričić L, Leskovac M. Micro Raman spec- Publisher’s Note Springer Nature remains neutral with regard to
troscopy of silica nanoparticles treated with aminopropylsilane- jurisdictional claims in published maps and institutional affiliations.
triol. Macromol Symp. 2008;265:178–82. https:​ //doi.org/10.1002/
masy.20085​0519.
41. Sydorchuk V, Khalameida S, Zazhigalov V, Skubiszewska-Zięba J,
Leboda R, Wieczorek-Ciurowa K. Influence of mechanochemical

13

Potrebbero piacerti anche