You are on page 1of 9

Research Article

Received: 4 November 2015 Revised: 15 February 2016 Accepted article published: 4 March 2016 Published online in Wiley Online Library: 5 April 2016

(wileyonlinelibrary.com) DOI 10.1002/jctb.4943

Enhanced photocatalytic degradation


of 4-chlorophenol and 2,4-dichlorophenol on
in situ phosphated sol-gel TiO2
Claudia Castañeda,a* Francisco Tzompantzi,a Ricardo Gómeza
and Hugo Rojasb

Abstract
BACKGROUND: Phosphated TiO2 photocatalysts with different phosphate content were synthesized by in situ phosphatation of
sol-gel TiO2 . H3 PO4 was used as a hydrolysis catalyst and as an anionic precursor. The solids were analyzed by several techniques
to determine the structural, textural and optical characteristics. The photoactivity of the materials was tested in the degradation
of 4-chlorophenol and 2,4-dichlorophenol, under UV irradiation.

RESULTS: XRD patterns showed that the presence of phosphate anions decreases the crystallite size without modification
of the anatase phase. Specific surface areas higher than the pristine TiO2 were obtained on the phosphated samples. The
presence of bidentate phosphates attached to the surface was identified by FTIR. The photoactivity of the materials tested in the
degradation of 4-chloropehol and 2,4-dichlorophenol showed notable improvement on phosphated TiO2 . The degradation rate
of the chlorophenolic compounds increases with the degree of substitution in the aromatic ring. Thus, the degradation rate of
2,4-dichlorophenol is higher than that of 4-chlorophenol. High efficiency in the degradation of the molecules after the recycling
of the photocatalysts was observed.

CONCLUSION: The modification in situ of TiO2 with low phosphate content promotes higher photocatalytic activity than that of
unmodified TiO2 and Degussa P-25 in the photodegradation of the chlorophenolic compounds.
© 2016 Society of Chemical Industry

Keywords: phosphated TiO2 ; sol-gel method; 4-chlorophenol photo-oxidation; 2,4-dichlorophenol photo- oxidation; photocatalysis

INTRODUCTION the characteristics of this oxide. Ortiz-Islas et al.17 reported that


Chlorophenols are widely used in several industrial processes, such phosphate anions stabilize the titania in the anatase phase and
as fabrication of pesticides, insecticides and wood preservatives. favor the development of Lewis acid sites. Connor et al.18 found
Since they have half-life times in water from 0.6 to 550 h and in that the phosphates bind strongly on TiO2 as bidentate surface
sediments up to 1700 h, they can be found in wastewater, ground- species. Two species of orthophosphates adsorbed on titania
water or soils.1,2 Advanced oxidation processes, among them pho- were reported by Gong:19 bridging bidentate surface complex
tocatalysis, have been used in recent decades for the degrada- (TiO)2 PO2 with C2v symmetry and electrostatically adsorbed PO4 −3
tion of highly persistent and non-biodegradable pollutants in the ion with Td symmetry. Goswami et al.20 synthesized nanocrys-
treatment of wastewater. talline phosphate doped TiO2 by the hydrothermal method and
Titanium dioxide (TiO2 ) is considered a promising semiconduc- they report that the incorporation of phosphates stabilized the
tor in heterogeneous catalysis due to its interesting chemical crystalline phase and caused a significant change in the TiO2 parti-
and physical properties. Its applications include use as catalytic cle size and in its textural properties, which resulted in an increase
supports,3 photocatalysts,4,5 in the production of hydrogen,6,7 in
solar cell electrodes8,9 and gas sensors.10,11 As a photocatalyst, its
catalytic behavior depends on the particle size, surface area, poros- ∗ Correspondence to: C Castañeda, Universidad Autónoma Metropolitana-
ity, the presence of surface acid or basic groups, light intensity and Iztapalapa, Depto. de Química, Área de Catálisis, Grupo ECOCATAL, Av. San
the nature of materials to be degraded.12,13 Rafael Atlixco No 186, México 09340, D.F. México.
Email: claudia.castaneda.mar@gmail.com
Modifications of TiO2 with anions have been reported to dimin-
ish the electron–hole recombination on the surface of titanium a Universidad Autónoma Metropolitana-Iztapalapa, Depto. de Química, Área de
dioxide.14 Thus, it has been reported that surface modification Catálisis, Grupo ECOCATAL, Av. San Rafael Atlixco No 186, México 09340, D.F.
of TiO2 by SO4 2− changes the band gap energy, increases the México
surface area, decreases the crystallite size and improves the pho- b Universidad Pedagógica y Tecnológica de Colombia, Escuela de Ciencias
tocatalytic activity.15,16 On the other hand, the phosphatation of Química, Grupo de Catálisis-UPTC, Avenida Central de Norte, Vía Paipa, Tunja,
2170

TiO2 has been reported as an effective method for enhancing Boyacá, Colombia

J Chem Technol Biotechnol 2016; 91: 2170–2178 www.soci.org © 2016 Society of Chemical Industry
Enhanced photodegradation of 4-CP and 2,4-DCP on in situ phosphated sol-gel TiO2 www.soci.org

in the photodegradation of fluorescein dye. Yu et al.21 reported average crystallite size of powders (D) was determined from the
that phosphated mesoporous TiO2 presents higher photoactivity XRD patterns, according to the Scherrer equation (D = K 𝜆 / 𝛽
than pure TiO2 and commercial P-25 in the oxidation of n-pentane. cos 𝜃), where K is a constant (shape factor, about 0.9), 𝜆 the
They attributed the improvement in activity to the extended band X-ray wavelength (1.5405 nm), 𝛽 the full width at half maximum
gap energy, increased surface area and the presence of Ti ions in of the diffraction peak and 𝜃 the diffraction angle.23 The values
tetrahedral coordination. In contrast, Colón et al.22 modified TiO2 of 𝛽 and 𝜃 of anatase phase were taken from (1 0 1) diffrac-
with different acid pre-treatments (nitric, sulfuric and phosphoric tion planes. N2 adsorption-desorption isotherms at –196∘ C were
acid). They found that the photoactivity in the degradation of phe- measured in a Quantachrome Autosorb 3B. Samples were previ-
nol decreased after treatment with phosphoric acid. This behavior ously evacuated at 250∘ C for 6 h. The Brunauer–Emmett–Teller
was attributed to the formation of pyrophosphate species on the (BET) method was used to calculate the total surface areas of
surface of the material. the samples and the pore size distribution was determined using
With this in mind, to study the effect of phosphates on the pho- the Brunauer–Joyner–Hallenda (BJH) method applied to the des-
tocatalytic properties of TiO2 , in the present work, phosphated orption branch. Diffuse reflectance spectra were obtained with a
TiO2 semiconductors were prepared by in situ phosphatation Cary 100 UV–Vis spectrophotometer equipped with an integrating
of sol–gel TiO2 and they were tested in the photodegradation sphere. The band gap energy values were calculated considering
of 4-chlorophenol (4-CP) and 2,4-dichlorophenol (2,4-DCP), two the indirect allowed transitions of the TiO2 semiconductor from
well-known pollutants. The phosphate anion was chosen because the plot of the modified Kubelka–Munk function [F(R) × h𝜈]1/2 ver-
of its strong adsorption on the TiO2 surface, its large negative sus energy of absorbed light.24 The FTIR absorption spectra of the
charge and its propensity to form hydrogen bonds. The photocat- samples were obtained with a Shimadzu IRAffinity-1 FTIR spec-
alysts were prepared varying the phosphate content (0.5, 1, 3 and trometer in the wavenumber range 4000–350 cm−1 .
5 wt%). In addition, recycling of the best photocatalysts was per-
formed after each test to assess the stability of the phosphates on Photocatalytic degradation
the semiconductor surface. To determine the optimum quantity of photocatalyst to per-
form the photocatalytic tests, degradations of 4-chloropenol (0.31
mmol L−1 ) using PTi1 sample were studied employing different
EXPERIMENTAL masses of semiconductor (25, 50, 100 and 200 mg). Study at three
Photocatalysts synthesis calcination temperatures (300∘ C, 400∘ C and 500∘ C, 0.5∘ C min−1 , 12
Phosphated titanium dioxide (PTi) was synthesized in situ via h) was performed with the optimum quantity of catalyst.
the conventional sol–gel route. Titanium butoxide (Aldrich 98%) Different solutions were prepared with molar concentra-
and butyl alcohol (Aldrich 99.4%) were mixed in a glass flask. tions of 0.31 mmol L−1 4-chlorophenol (Sigma 99%) and
Phosphoric acid (H3 PO4 , Baker Analyzed 84.8%) was used as a 2,4-dichlorophenol (Aldrich 99%), respectively. For each test,
hydrolysis catalyst and as an anionic precursor. H3 PO4 was added 50 mg of the calcined photocatalyst were added to 200 mL
in different amounts to the above mixture in order to obtain of aqueous solution in a glass reactor. The suspension was
TiO2 with different phosphate content (0.5, 1, 3 and 5% wt). The stirred at 700 rpm in the dark for 30 min with air flow, to ensure
mixture was maintained at 40∘ C under vigorous stirring for 2 h. adsorption–desorption equilibrium of the test molecule. Subse-
Distilled water was added to 50 mL of ethanol and the solution was quently, the solution was irradiated with UV light using a Pen Ray
added dropwise to the reaction mixture to initiate the hydrolysis Power Supply lamp, 𝜆 = 254 nm, 4.4 mW protected by a quartz
process. The molar ratio of titanium butoxide:water:butyl alcohol tube, immersed in the solution. Degradation of the pollutants
was 1:8:35, respectively. The resulting mixture was refluxed at 75∘ C was monitored by analyzing an aliquot sample every 30 min and
for 24 h under vigorous stirring. After, the gel obtained was rinsed following the absorption band of the test molecule (278 nm for
with distilled water and dried at 85∘ C for 24 h (Xerogel). Finally, 4-chlorophenol and 283 nm for 2,4-dichlorophenol) using a UV-Vis
the powder was calcined at 400∘ C for 12 h with a heating ramp spectrophotometer Cary 100. Total organic carbon (TOC) was also
of 0.5∘ C per min. The calcination temperature was chosen from quantified (Shimadzu TOC-V CSN).
previous study of the optimal calcination conditions. The solids
were denoted as PTi0.5, PTi1, PTi3 and PTi5 corresponding the digit
with the phosphates percentages. For comparison, pure TiO2 was RESULTS AND DISCUSSION
synthesized using the same procedure with the addition of nitric Characterization of the photocatalysts
acid instead of H3 PO4 as hydrolysis catalyst. In addition, in some Energy dispersion X-ray spectroscopy
cases, a commercial TiO2 (Degussa P-25) was used for comparative Semi-quantitative analysis of the phosphorus content using EDX
behavior. was made for the PTi1 and PTi5 photocatalysts. For comparison,
EDX analysis for TiO2 was performed. The spectra of the samples
studied are presented in Fig. 1, where the characteristic signals for
Characterization of photocatalysts titanium (Ti) and oxygen (O) at 4.50 and 0.52 keV, respectively, can
The semi-quantitative analysis of the phosphorus content in the be observed. The distinctive signal for phosphorus (2.01 keV) is
calcined photocatalysts was performed in a scanning electron present for both phosphated samples and its intensity is greater
microscope JEOL 7600F, equipped with an energy dispersion X-ray for PTi5.
(EDX) spectrometer from Oxford instruments. From this informa- From these results, the phosphate content was estimated at
tion, determination of the phosphate content in the materials was 1.65% and 6.45% for the PTi1 and the PTi5 samples, respectively.
obtained. XRD patterns and phase identification of the samples The anion percentages are higher in the calcined material due to
were collected on a Bruker D2 Phaser diffractometer using Cu the loss of water and alcohol during activation at 400∘ C. Similar
K𝛼 radiation. The samples were scanned in the 2𝜃 range 20–80∘ results were reported by Pattnayak and Parida25 and Samantaray
2171

with a step size of 0.02∘ and a counting time of 0.6∘ sec−1 . The and Parida.26

J Chem Technol Biotechnol 2016; 91: 2170–2178 © 2016 Society of Chemical Industry wileyonlinelibrary.com/jctb
www.soci.org C Castañeda et al.

Figure 1. EDX spectra of the TiO2 , PTi1 and PTi5 samples, calcined at 400∘ C.

(101)
(004) (200) (105) (204) Table 1. Textural and structural parameters of the calcined photo-
PTi5 catalysts at 400∘ C
Intensity (a.u.)

PTi3
Surface Pore
PTi1
Eg area volume Crystallite
Sample (eV) (m2 g−1 ) Pore size (Å) (cm3 g−1 ) size (nm)
PTi0.5

TiO2 TiO2 3.01 126 56 0.23 10.1


PTi 0.5% 3.08 185 34 0.23 7.3
20 30 40 50 60 70 80 PTi 1% 3.11 198 34 0.22 7.1
2 (degrees) PTi 3% 3.09 269 34 0.24 5.6
PTi 5% 3.10 285 34 0.26 5.8
Figure 2. X-ray diffraction patterns for the TiO2 and phosphated TiO2
samples, calcined at 400∘ C.

Nitrogen adsorption
Powder X-ray diffraction Figure 3(A) shows the adsorption–desorption isotherms for the
The diffraction patterns of the phosphated TiO2 and unmodified modified samples and reference TiO2 calcined at 400∘ C. All the
TiO2 are shown in Fig. 2. The materials calcined at 400∘ C exhibit samples exhibited the type IV nitrogen isotherm, which is char-
peaks located at 2𝜃 angles of 25.4∘ (101), 37.9∘ (004), 48.1∘ (200), acteristic of mesoporous materials.27 The H2 type hysteresis was
54.5∘ (105), 62.8∘ (204) and 75.2∘ (215), which correspond to the observed in the four samples which are often referred to as
anatase phase (JCPDS 84-1286). The rutile phase was not observed an ‘ink-bottle’ pore. The existence of the hysteresis loop in the
in any of the phosphated and reference sol–gel TiO2 samples. This isotherm is due to the capillary condensation of nitrogen gas
indicates that the phosphate does not modify the TiO2 crystalline occurring in the mesoporous cavity.28 The phosphatation in situ
phase. of the TiO2 semiconductor induces a displacement toward lower
Upon increasing the phosphate content, the peak intensity for relative pressures in the desorption process, suggesting a change
anatase (2𝜃 = 25.4∘ ) gradually decreases and the half-width of in the mesoporosity of the material. The BET specific surface area,
the diffraction peak is broadened, indicating the formation of pore size and pore volume of the phosphated samples and the
smaller TiO2 crystallites. The average crystallite size was calculated unmodified TiO2 are listed in Table 1. The modification of TiO2
using the Scherrer equation from the broadening of this peak, with phosphates is beneficial to obtain solids with larger sur-
corresponding to the plane (101) (Table 1). As can be observed, face areas, which could be related with the hydroxylation extent
the crystallite size became smaller as the phosphate content of the material generated during the gelification of the titanium
increased. This result suggests that the phosphatation inhibits the butoxide in presence of phosphoric acid. The monomodal pore
2172

growth of the TiO2 crystallites. size distribution (Fig. 3(B)), derived from the desorption branch of

wileyonlinelibrary.com/jctb © 2016 Society of Chemical Industry J Chem Technol Biotechnol 2016; 91: 2170–2178
Enhanced photodegradation of 4-CP and 2,4-DCP on in situ phosphated sol-gel TiO2 www.soci.org

(A) (B)
3.5

Desorption Dv (logd)/(cm3/g)
PTi5 TiO2

Adsorbed Quantity (a.u.)


PTi3 3.0
PTi0.5
2.5 PTi1
PTi1
PTi3
2.0 PTi5
PTi0.5
1.5
TiO2
1.0
0.5
0.0
0.0 0.2 0.4 0.6 0.8 1.0 20 30 40 50 60 70 80
P/Po Pore Size (Å)

Figure 3. (A) Nitrogen adsorption–desorption isotherms, and (B) pore size distribution for the TiO2 and phosphated TiO2 samples, calcined at 400∘ C.

the isotherms indicates a decrease in the pore size of the phos- (A)
phated materials in relation to TiO2 ; however this property does 1.6 269 330 TiO2
not change with the increase of the anion content. Thus, the 225
1.4 PTi0.5
phosphatation process probably promotes a pore blocking, which 1.2 PTi1
PTi3
causes a change in the mesoporosity of the TiO2 semiconductor. 1.0 PTi5

F(R)
0.8
UV-Vis absorption and diffuse reflectance spectroscopy 0.6
411
To observe the optical response of the phosphated TiO2 , the UV-Vis 0.4
absorption spectra were recorded in a wavelength range between 0.2
190 and 500 nm (Fig. 4(A)). A strong absorption band between 300
0.0
and 400 nm was observed. This absorption band is characteristic 200 300 400 500
of O2p → Ti3d transitions in the tetrahedral symmetry. The two
Wavelength (nm)
shoulders that appear from approximately 190–240 and 250–300
nm, present in both the unmodified TiO2 as well as the phosphate (B)
3.0
TiO2 , may be associated with bulk crystals.24
Unmodified TiO2 and phosphated materials show clear ultra- 2.5
violet absorption. The unmodified TiO2 powder exhibits an 2.0
1/2

absorption band edge at 411 nm. However, phosphated samples


[F(R) x hv]

present a blue-shifted in comparison with the optical response of 1.5


unmodified TiO2 . This behavior is attributed to the well-known 1.0 TiO2
quantum-size effect for semiconductors. UV-visible absorption PTi0.5
band edge is a direct function of the crystallite size for diameters 0.5 PTi1
PTi3
less than 10 nm.12 The increase in the phosphate content causes 0.0 PTi5
a decrease in the crystallite size and hence an absorption toward
2 3 4 5 6
lower wavelengths. The band gap energy values (Eg) calculated
from the modified Kubelka–Munk plot (Fig. 4(B)) were obtained Energy (eV)
considering indirect transitions of the semiconductors (reported Figure 4. (A) UV-Vis spectra for the TiO2 and phosphated TiO2 sam-
in Table 1). The estimated band gap energy for TiO2 is 3.01 eV ples, calcined at 400∘ C, and (B) Kubelka–Munk modified spectra for the
and a slight increase between 3.08 and 3.11 eV was observed as a photocatalysts.
result of the phosphatation of the material.
was detected at 1670 cm−1 .30 In the modified photocatalysts, these
Fourier transform infrared spectroscopy latter signals are broader, indicating that the phosphatation of
FTIR spectra of the phosphated TiO2 photocatalysts were used to TiO2 increases the amount of surface-adsorbed water and hydroxyl
analyze the structure of the materials (Fig. 5(A)–(C)). In the region groups. This behavior is in agreement with the larger surface
of low energy 500–722 cm−1 the stretching vibrations of Ti–O areas observed for the phosphated samples. A band at 1050 cm−1
and Ti–O–Ti bonds, typical of titanium dioxide,29 can be seen in appears in the phosphated TiO2 materials and is attributed to the
Fig. 5(A). phosphate in a bidentate state of ligation; 31 the band at 2329
An amplification of the infrared spectrum was performed in cm−1 is assigned to vibration of the P–OH bond which increase
order to elucidate better the vibration bands (Fig. 5(B)). As can be proportionally with the phosphate content.32 A signal at 1125
observed, a wide band at 3243 cm−1 corresponding to adsorbed cm−1 associated with a Ti–O–P bond, with P replacing Ti4+ in the
water (Ti-OH2 ) is observed in all samples. The signal at 1510 cm−1 TiO2 lattice, was not observed.33
is assigned to the combination of the symmetric and asymmetric The infrared spectra of the photocatalysts after use in the
stretching modes of water molecules adsorbed on the TiO2 surface degradation reactions can be observed in Fig 5(C). The vibration
2173

and the band related with the flexion vibration of H–O groups bands at 1050 cm−1 attributed to the phosphates are still present;

J Chem Technol Biotechnol 2016; 91: 2170–2178 © 2016 Society of Chemical Industry wileyonlinelibrary.com/jctb
www.soci.org C Castañeda et al.

(A) (B)
1050
PTi5 3243 2329 1670
PTi5 1510
PTi3

Absorbance
Absorbance
PTi1 PTi3
PTi1
PhTi0.5
TiO 2 PTi0.5
TiO
2

3500 3000 2500 2000 1500 1000 500 3500 3000 2500 2000 1500 1000
Wavenumber (cm-1) Wavenumber (cm-1)

(C)
1050
2329
PTi5
Absorbance PTi3

PTi1

PTi0.5

TiO 2

3500 3000 2500 2000 1500 1000


Wavenumber (cm-1)

Figure 5. (A) FTIR spectra for the TiO2 and phosphated TiO2 samples, calcined at 400∘ C, (B) enlargement of the infrared spectrum, and (C) FTIR spectra for
used photocatalysts in the degradation reactions.

confirming the stability of the phosphates bonded to the surface were calcined at three temperatures (300∘ C, 400∘ C and 500∘ C)
of TiO2 during the photodegradation reactions. As expected, the and tested in the photoreaction during 3 h. As can be seen in
bands at 1510 and 1670 cm−1 were more intense due to the hydra- Fig. 6(B), degradation of the contaminant using the PTi1 sample
tion suffered by the solid in contact with the reactant aqueous calcined at 300∘ C is not efficient; however, when the reaction is
solutions. carried out with the semiconductor calcined at 400∘ C, the photo-
catalytic activity dramatically improves and a degradation of 68%
was attained after 210 min. Increasing further the calcination tem-
Photocatalytic degradation
perature to 500∘ C notably the photoactivity was suppressed and
Optimization of the photocatalyst quantity and of the calcination
temperature for the photocatalytic test only 29% of the molecules were oxidized.
Further photoactivity tests of the samples were performed using
In order to determine the optimal mass of solid to perform the
50 mg of the semiconductor calcined at 400∘ C in the degradation
photocatalytic test, the degradation of 200 mL of a solution of
of 4-clorophenol and 2,4-diclorophenol.
4-clorophenol was performed using different quantities of PTi1
(25, 50, 100 and 200 mg). In Fig. 6(A) it can be seen that 50 mg
of semiconductor were sufficient to degrade the contaminant; at Degradation of 4-cholophenol
higher solid contents the reaction remains constant and diffusion The photodegradation capacities of the phosphated titania
effects, such as the light intensity and mass transfer diffusion are were tested with two important pollutants, 4-chlorophenol
evident. and 2,4-dichlorophenol (2,4-DCP). The contaminant concen-
On the other hand, to determinate the most favorable calcina- tration was monitored by UV-Vis spectroscopy applying the
tion temperature for the photodegradation of 4-CP, 50 mg of PTi1 Beer–Lambert law.

(A) (B)
70
4-CP photodegradation (%)

1.8
1.7 60
K x 10-6 (M min-1)

1.6 50
1.5 40
1.4 30
1.3 20
1.2 10
1.1 0
25 50 75 100 125 150 175 200 300 400 500
Mass Photocatalyst (mg) Calcination temperature (°C)
2174

Figure 6. (A) Effect of the mass, and (B) effect of the calcination temperature of the PTI1 photocatalyst in the photodegradation of 4-CP after 3 h reaction.

wileyonlinelibrary.com/jctb © 2016 Society of Chemical Industry J Chem Technol Biotechnol 2016; 91: 2170–2178
Enhanced photodegradation of 4-CP and 2,4-DCP on in situ phosphated sol-gel TiO2 www.soci.org

3.5 0min
Adsorption
30min
3.0 60min
90min
2.5

Absorbance
120min
150min
2.0 180min

1.5
1.0 283
0.5
0.0

200 250 300 350


Figure 7. UV absorption spectra of the degradation of 4-CP using the PTi1 Wavelength (nm)
sample.
Figure 9. UV absorption spectra of the degradation of 2,4-DCP using the
PTi1sample.
70
4-CP photodegradation (%)

60
50 90

2,4-DCP photodegradation (%)


40
80
70
30
60
20
50
10
40
0% 0%
0 30
TiO2 PTi0.5 PTi1 PTi3 PTi5 P-25
20
Figure 8. Photocatalytic degradation of 4-CP for the studied photocatalysts
after 3 h reaction. 10
0
TiO2 PTi0.5 PTi1 PTi3 PTi5 P-25

Figure 7 illustrates the changes in the UV absorption spectra of Figure 10. Photocatalytic degradation of 2,4-DCP for the studied photo-
4-CP at different irradiation times. 4-chlorophenol presents strong catalysts after 3 h reaction.
absorption bands in the ultraviolet region with wavelengths at
195, 223 and 278 nm. As the reaction progresses, the characteristic
peaks of 4-CP at 𝜆 = 278 nm and 𝜆 = 223 nm decreased in height.
An increase in the absorbance is observed near 250 nm and
can be attributed to the formation of intermediates, especially O
benzoquinones. 34 In addition, the absorption spectrum displays O
a bathochromic shift towards 285 nm, indicating the formation of P
O
intermediary species with a different cromophore group, such as O
hydroquinone.35 This last peak (285 nm), assigned to the n → 𝜋*
transitions of C–Cl bond, decreases rapidly because this bond
is the most reactive. Finally, the absorption bands located near
190 and 230 nm are assigned to the 𝜋 → 𝜋* transitions of the
aromatic group.36 These bands present a significant decrease with
Ti

irradiation time.
As can be seen in Fig. 8, the TiO2 , PTi0.5, PTi1 photocatalysts and
Degussa P-25 degraded in nearly 45, 62, 68 and 46%, respectively.
It should be noted that the activity of the phosphated sample at
1% is higher than that of the reference sol–gel TiO2 and P-25 in
the degradation of 4-CP. Probably the small crystallite size present
in PTi1 facilitates surface charge transfer, which decreases the
recombination of electron–hole pairs and improves the pho-
tocatalytic activity.37 However, at high phosphate content the
photoactivity was inhibited. Although the PTi3 and PTi5 photo-
catalysts present larger surface areas and a smaller average crystal
size, the elevated PO4 3− content could block the active sites on Figure 11. Illustrative design explaining the role of phosphates in
the TiO2 surface diminishing the interaction of the pollutant with chlorophenols photodegradation.
2175

the semiconductor.

J Chem Technol Biotechnol 2016; 91: 2170–2178 © 2016 Society of Chemical Industry wileyonlinelibrary.com/jctb
www.soci.org C Castañeda et al.

(A) (B)
80 40
70

Residual TOC (%)

Residual TOC (%)


60 30
50
40 20
30
20 10
10
0 0
TiO2 PTi1 P-25 TiO2 PTi1 P-25

Figure 12. Residual TOC in the photodegradation of (A) 4-CP and (B) 2,4-DCP using TiO2 , P-25 and PTi1 samples after 3 h reaction.

(A) (B)
0.0004
1.45 x 10-6 M min-1 0.0003
1.60 x 10-6 M min-1
Concentration [M]

Concentration [M]
1.27 x 10-6 M min-1
0.0003 1.59 x 10-6 M min-1
0.0002

0.0002
PTi1 PTi1
TiO2 0.0001
TiO2
0.0001 P-25 1.77 x 10-6 M min-1 P-25 1.87 x 10-6 M min-1

0 30 60 90 120 150 0 30 60 90 120 150


Time (min) Time (min)
Figure 13. Zero-order kinetics for the photodegradation of (A) 4-CP and (B) 2,4-DCP using the PTi1sample.

90 more slowly, which indicates that the breakage of C–Cl bonds is


Cycle 1
80 Cycle 2
preferred.
Photodegradation (%)

70 Figure 10 presents the 2,4-DCP photodegradation percentages


obtained by the studied photocatalysts. In the presence of TiO2
60
and phosphated photocatalysts (0.5, 1, 3 and 5%), oxidations of
50 70, 77, 83, 62 and 39% were achieved, respectively, in the same
40 reaction time. When the test was performed with Degussa P-25,
30 nearly 71% of molecules were degraded.
20 Similar to that observed in 4-chlorophenol photodegradation,
a low level of phosphates, particularly the PTi1 solid shows the
10
higher photocatalytic activity than reference sol–gel TiO2 and
0
4-CP 2,4-DCP Degussa P-25. On the other hand, an increase in the phos-
phate content above 3% negatively affects the activity. In a
Figure 14. Recycling of the PTi1 photocatalyst for the degradation of 4-CP
similar manner to that proposed for 4-chlorophenol, the high
and 2,4-DCP.
anion content on the surface could block the substrate adsorp-
tion and transfer between the hole and the substrate can be
hindered.
Degradation of 2,4-dichlorophenol From the results obtained in this study, the phosphates anchored
The UV absorption spectra of 2,4-DCP degradation at different to the surface of TiO2 could obstruct the recombination of holes
irradiation times using PTi1 is shown in Fig. 9. The absorption and electrons. The negative charge of the phosphates generates a
band at 283 nm is associated with the 2,4-DCP. When the pol- positive partial charge on titanium which attracts the photogener-
lutant is placed in contact with the photocatalyst, without UV ated electrons to the surface. The electrons reduce the oxygen to
irradiation, a decrease in the absorbance is observed. This behav- form superoxide radicals and the photogenerated holes migrate
ior implies strong adsorption of the organic compound on the toward the surface to oxidize the organic pollutant. An illustration
surface of the solid. When the suspension is irradiated, an increase of the role of the phosphates on the surface of TiO2 is shown in
in the absorbance is evidenced. This could be indicating release Fig. 11.
of the pollutant from the surface of the photocatalyst. As the Determination of the mineralization of the organic compounds
photocatalytic reaction progresses, the absorption band at 283 was evaluated by measuring the total organic carbon (TOC) of the
nm (n → 𝜋* transitions) gradually decreases, which is a result of irradiated solution. Figure 12(A) and (B), respectively, presents the
the photo-oxidation of the compound. The 203 nm band asso- residual TOC results obtained after 3 h reaction, in the degradation
2176

ciated with 𝜋 → 𝜋* transitions of the aromatic ring disappears of 4-CP and 2,4-DCP using TiO2 , PTi1 photocatalysts and Degussa

wileyonlinelibrary.com/jctb © 2016 Society of Chemical Industry J Chem Technol Biotechnol 2016; 91: 2170–2178
Enhanced photodegradation of 4-CP and 2,4-DCP on in situ phosphated sol-gel TiO2 www.soci.org

P-25. The TOC measurements confirm that the PTi1 sample is more ACKNOWLEDGEMENT
efficient than both TiO2 and Degussa P-25 for mineralization of the The authors are grateful to CONACYT for the scholarship num-
organic pollutants. Comparing the behavior of the two molecules, ber 287123 and for the financial support grants 194451 Materiales
it can be seen that the oxidation of 2,4-DCP was greater than that nanoestructurados para fotocatálisis and FOINS/75/2012 Fotosín-
obtained for 4-CP. This result indicates that the total mineralization tesis artificial.
of chlorophenols is greater with more chlorine substitution in the
aromatic ring.
From the photoactivity data as a function of time for the PTi1 REFERENCES
sample, TiO2 and Degussa P-25, the kinetic behaviors of the pho- 1 Andreozzi R, Di Somma I, Marotta R, Pinto G, Pollio A and Spasiano D,
todegradation reactions were analyzed. The photodegradations of Oxidation of 2,4-dichlorophenol and 3,4-dichlorophenol by means
4-CP and 2,4-DCP follow zero-order kinetics, represented in Fig. 13. of Fe(III)-homogeneous photocatalysis and algal toxicity assess-
ment of the treated solutions. Water Res 45:2038–2048 (2011).
The 4-CP and 2,4-DCP degradation rates were higher with the 2 Eberlin MN and Cesar da Silva R, Faster and simpler determination
phosphated photocatalyst PTi1 than with TiO2 and Degussa P-25. chlorophenols in water by fiber introduction mass spectrometry.
The apparent rate constant for the degradation of chlorophenols Anal Chim Acta 620:97–102 (2008).
increases with the number of chlorides in the molecule. This obser- 3 Rooke JC, Barakat T, Finol MF, Billemont P, De Weireld G, Li Y et al.,
Influence of hierarchically porous niobium doped TiO2 supports
vation is in agreement with the results reported by Hugül et al.38 in in the total catalytic oxidation of model VOCs over noble metal
the decomposition of 2-CP, 2,4-DCP and 2,4,6-TCP by UV/hydrogen nanoparticles. Appl Catal B 142–143:149–160 (2013).
peroxide oxidation. A similar kinetic order was reported by 4 Min Y, Zhang K, Chen Y and Zhang Y, Synthesis of novel visible
Sivalingam et al.39 in the photodegradation of substituted phenols light responding vanadate/TiO2 heterostructure photocatalysts for
application of organic pollutants. Chem Eng J 175:76–83 (2011).
using solution combustion synthesized TiO2 (pentachlorophe-
5 Lopez T, Gomez R, Sanchez E, Tzompantzi F and Vera L, Photocatalytic
nol > trichlorophenol > dichlorophenol > p-chlorophenol). activity in the 2,4-dinitroaniline decomposition over TiO2 sol-gel
The behavior observed in this work can be explained by the derived catalysts. J Sol-Gel Sci Technol 22:99–107 (2001).
strong positive mesomeric effect (+M-effect) of the − OH group 6 Zhang J, Du P, Schneider J, Jarosz P and Eisenberg R, Photogeneration
of the aromatic ring, where it is expected that the major points of hydrogen from water using an integrated system based on
TiO2 and platinum(II) diimine dithiolate sensitizers. J Am Chem Soc
of attack by . OH radicals will be the positions of high electron 129:7726–7727 (2007).
density. Thus, para- and ortho- positions are the most active in 7 Fan L, Long J, Gu Q, Huang H, Lin H and Wang X, Single-site
photodegradation of chlorophenols.40 For this reason, 2,4-DCP nickel-grafted anatase TiO2 for hydrogen production: toward
was more vulnerable to hydroxyl radical attack, evidenced by the understanding the nature of visible-light photocatalysis. J Catal
320:147–159 (2014).
greater reactivity compared with that of 4-chlorophenol. 8 Bwana N, Effects of the morphology of the electrode nanostruc-
An important parameter to assess the application of photocat- tures on the performance of dye-sensitized solar cells. Nano Res
alysts in the degradation of organic pollutants is the recycling 1:483–489 (2008).
experiment. Favorably, PTi1 was recycled after being recovered 9 Motlak M, Barakat NAM, Akhtar MS, Hamza AM, Yousef A, Fouad H et
al., Influence of GO incorporation in TiO2 nanofibers on the elec-
from the solution and subjected to heat treatment at 400∘ C for 3 h.
trode efficiency in dye-sensitized solar cells. Cera Int 41:1205–1212
This experiment was made three times to analyze the reproducibil- (2015).
ity and the statistical values obtained are reported as error bars in 10 Landau O and Rothschild A, Microstructure evolution of TiO2
Fig. 14. The photocatalytic behavior of the calcined and recycled gas sensors produced by electrospinning. Sensors Actuators B
samples in the degradation of 4-CP and 2,4-DCP after 3 h reaction 171–172:118–126 (2012).
11 Seo MH, Yuasa M, Kida T, Huh JS, Yamazoe N and Shimanoe K,
is shown in Fig. 14. As can be observed, the photocatalytic degra- Detection of organic gases using TiO2 nanotube-based gas sensors.
dation of 4-CP decreased from 68% to 64% and the oxidation of Procedia Chem 1:192–195 (2009).
2,4-DCP decreased from 83% to 80%. The results confirm that PTi1 12 Zhang Q, Gao L and Guo J, Effects of calcination on the photocatalytic
can be employed as a promising reusable material for removal of properties of nanosized TiO2 powders prepared by TiCl4 hydrolysis.
Appl Catal B 26:207–215 (2000).
chlorophenols. 13 Bacsa RR and Kiwi J, Effect of rutile phase on the photocatalytic
properties of nanocrystalline titania during the degradation of
p-coumaric acid. Appl Catal B 16:19–29 (1998).
14 Kőrösi L, Oszkó A, Galbács G, Richardt A, Zöllmer V and Dékány
CONCLUSIONS I, Structural properties and photocatalytic behaviour of
phosphate-modified nanocrystalline titania films. Appl Catal B
The in situ modification of TiO2 with phosphate anions promotes 77:175–183 (2007).
the formation of solids with high specific surface areas and small 15 Samantaray SK and Parida KM, Effect of anions on the textural and
crystallite size. PTi1 sample was more photoactive than reference catalytic activity of titania. J Mater Sci 38:1835–1848 (2003).
16 Muggli DS and Ding L, Photocatalytic performance of sulfated TiO2
sol–gel TiO2 and Degussa P-25 in the degradations of 4-CP and
and Degussa P-25 TiO2 during oxidation of organics. Appl Catal B
2,4-DCP. On samples with low phosphate content (PTi 1 wt%), a 32:181–194 (2001).
favorable degradation of organic pollutants such as chlorophenols 17 Ortiz-Islas E, López T, Gómez R and Navarrete J, Effect of phos-
was obtained, while for samples with phosphate content of 3 phate ions in the properties of titania sol-gel. J Sol–Gel Sci Technol
wt% or higher a possible blockage of the TiO2 surface by the 37:165–168 (2006).
18 Connor PA and McQuillan AJ, Phosphate adsorption onto TiO2 from
phosphate anions could cause a decrease in the photoactivity. aqueous solutions: an in situ internal reflection infrared spec-
The degradation rate increases with substitution of chlorine into troscopy study. Langmuir 15:2916–2921 (1999).
the phenolic ring; therefore, the oxidation of 2,4-dichlorophenol 19 Gong W, A real time in situ ATR-FTIR spectroscopy study of linear
is greater than that of 4-chlorophenol. The photocatalysts can be phosphate adsorption on titania surfaces. Int J Minerals Process
63:147–165 (2001).
easily recycled and the phosphates bound to the surface of TiO2 20 Goswami P and Ganguli JN, Synthesis, characterization and photocat-
are stable, resulting in an acceptable degree of activity retention alytic reactions of phosphated mesoporous titania. Bull Mater Sci
2177

upon reuse. 35:889–896 (2012).

J Chem Technol Biotechnol 2016; 91: 2170–2178 © 2016 Society of Chemical Industry wileyonlinelibrary.com/jctb
www.soci.org C Castañeda et al.

21 Yu JC, Zhang L, Zheng Z and Zhao J, Synthesis and characterization of 31 Ramadan AR, Yacoub N, Amin H and Ragai J, The effect of phosphate
phosphated mesoporous titanium dioxide with high photocatalytic anions on surface and acidic properties of TiO2 hydrolyzed from
activity. Chem Mater 15:2280–2286 (2003). titanium ethoxide. Colloids Surf A 352:118–125 (2009).
22 Colón G, Sánchez-España JM, Hidalgo MC and Navío JA, Effect of TiO2 32 Colthup NB, Daly LH and Wiberley SE, Compounds containing boron,
acidic pre-treatment on the photocatalytic properties for phenol silicon, phosphorous, sulfur,or halogen, in Introduction to Infrared
degradation. J Photochem Photobiol A: Chem 179:20–27 (2006). and Raman Spectroscopy, 3rd edn. Academic Press, San Diego
23 Ahmad M, Modified Scherrer equation to estimate more accurately (1990).
nano-crystallite size using XRD. World J Nano Sci Eng 2:154–160 33 Tran ATT, Duke M, Diniz Costa JCD and Gray PG, Proton conductivities
(2012). of titanium phosphate at high temperature for PEMFC. Asia Pacific J
24 López R and Gómez R, Photocatalytic degradation of 4-nitrophenol Chem Eng 14:101–118 (2006).
on well characterized sol–gel molybdenum doped titania semicon- 34 Thorsten Wilke MS and Kleinermanns K, 1,4-Hydroquinone is a hydro-
ductors. Top Catal 54:504–511 (2011). gen reservoir for fuel cells and recyclable via photocatalytic water
25 Pattnayak PK and Parida KM, Studies on PO4 3− /ZrO2 - II. Effect of phos- splitting. Open J Phys Chem 3:97–102 (2013).
phate concentration and activation temperature on the catalytic 35 Rychlinska I and Nowak S, Quantitative determination of arbutin
properties of zirconia. J Colloid Interface Sci 226:340–345 (2000). and hydroquinone in different plant materials by HPLC. Notulae
26 Samantaray SK and Parida K, Effect of phosphate ion on the textural Botanicae Horti Agrobotanici 40:109–113 (2012).
and catalytic activity of titania-silica mixed oxide. Appl Catal A Gen 36 Valente JS, Tzompantzi F and Prince J, Highly efficient photocatalytic
220:9–20 (2001). elimination of phenol and chlorinated phenols by CeO2 /MgAl lay-
27 Sing KSW, Everett DH, Haul RAW, Moscou L, Pierotti RA, Rouquerol J et ered double hydroxides. Appl Catal B 102:276–285 (2011).
al., Reporting physisorption data for gas/solid systems with special 37 Lin L, Lin W, Xie JL, Zhu YX, Zhao BY and Xie YC, Photocatalytic
reference to the determination of surface area and porosity. Pure properties of phosphor-doped titania nanoparticles. Appl Catal B
Appl Chem 57:603–619 (1985). 75:52–58 (2007).
28 Naumov S, Valiullin R, Käger J and Monson P, Understading adsorption 38 Hugül M, Apak R and Demirci S, Modeling the kinetics of UV/hydrogen
and desorption processes in mesoporous materials with indepen- peroxide oxidation of some mono-, di-, and trichlorophenols.
dent disordered chennels. Phys Rev E 80:1–9 (2009). J Hazard Mater 77:193–208 (2000).
29 López R, Gómez R and Oros S, Photophysical and photocatalytic 39 Sivalingam G, Priya MH and Madras G, Kinetics of the photodegra-
properties of TiO2 -Cr sol–gel prepared semiconductors. Catal Today dation of substituted phenols by solution combustion synthesized
166:159–165 (2011). TiO2. Appl Catal B 51:67–76 (2004).
30 Nakamoto K, Applications in coordination chemistry, in Infrared and 40 Moza PN, Fytianos K, Samanidou S and Korte F, Photodecomposition
Raman Spectra of Inorganic and Coordination Compounds, 6th edn. of chlorophenols in aqueous medium in presence of hydrogen
Wiley, New Jersey, (2008). peroxide. Bull Environ Contam Toxicol 41:678–682 (1988).
2178

wileyonlinelibrary.com/jctb © 2016 Society of Chemical Industry J Chem Technol Biotechnol 2016; 91: 2170–2178

You might also like