Journal of Physics and Chemistry of Solids: Sciencedirect

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Physics and Chemistry of Solids 134 (2019) 262–272

Contents lists available at ScienceDirect

Journal of Physics and Chemistry of Solids


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

Study on the effects of Cl and F doping in TiO2 powder synthesized by a sol- T


gel route for biomedical applications
Vijay Singha, Ankit Raoa, Anamika Tiwaria, Padarthi Yashwantha, Manohar Lalb, Uma Dubeyb,
Shampa Aichc, Banasri Roya,∗
a
Department of Chemical Engineering, BITS Pilani, Pilani Campus, Rajasthan, 333031, India
b
Department of Biotechnology, BITS Pilani, Pilani Campus, Rajasthan, 333031, India
c
Department of Metallurgical and Materials Engineering, Indian Institute of Technology, Kharagpur, West Bengal, 721302, India

ARTICLE INFO ABSTRACT

Keywords: F- and Cl-doped (10 at%) TiO2 powders were prepared by a sol-gel route and calcined at different temperatures
F-doped titanium oxide and for different times. X-ray diffraction, thermogravimetric- and differential-thermal analysis showed that
Cl-doped titanium oxide doping suppressed both the amorphous to anatase transformation and the anatase to rutile transformation and
Biocompatibility refines the particle size. F was found to be more effective than Cl. Preliminary hemolysis and cytotoxicity testing
X-ray photoelectron spectroscopy
demonstrated that F- and Cl-doped anatase powders were more hemolytic and toxic than the undoped anatase
Rietveld refinement
powder; however, the doped rutile powders were less hemolytic and toxic than the undoped rutile powder. The
lattice strain (calculated from Rietveld analysis) was less for the doped rutile powders, which could explain their
enhanced biocompatibility. X-ray photoelectron spectroscopy revealed the presence of Ti3+ in F-doped anatase
powder. Ti3+ ions create reactive oxygen species and made the doped anatase sample less biocompatible. X-ray
photoelectron spectroscopy revealed that when these doped samples were calcined at 400 °C or higher tem-
perature, the doping ions disappeared from the material. Absence of the dopants may generate a significant
defect per unit volume in anatase and influence the anatase to rutile nucleation density and particle size.
Presumably, these higher-energy defects present in the doped powders acted as new nucleation sites, and higher
temperature and/or a longer transformation time provided the ions with greater mobility to rearrange them-
selves. With a longer conversion time for the doped powders, the new rutile phase had less strain and greater
thermodynamic stability.

1. Introduction the biomedical field, especially as a hard tissue replacement. The


combination of the low specific weight and Young's modulus approxi-
Biomaterials, artificial or natural, are used to reinstate or replace an mately close to that of natural bone (high elastic modulus can result in
organ (fully or partially) to reclaim its structure and function so as to stress shielding) of these materials is viewed as a biomechanical ad-
improve the quality of human life and increase human longevity. vantage [7,8]. For good biocompatibility, high hardness, and high
Biomaterials can be used in many parts of the human body; for ex- corrosion and wear resistance of titanium-based metals/alloys, biome-
ample, as artificial valves in the heart, stents in blood vessels, a lens in dical professionals generally rely on the fabrication of a thin adherent
the eye, implants in hip, elbow, ear, shoulder, and knee, and dental oxide film on the surface. Also, it acts as a diffusion barrier against the
structures [1–3]. The current research target is to develop implants that transport of toxic metallic ions in biological systems.
can serve for very long periods or for a lifetime without failure or the Doping TiO2 with a suitable anion and/or cation is a very effective
necessity of revision surgery [4]. Thus, the development of material way to alter its chemical structure and surface properties and modify its
with high corrosion and wear resistance in the bodily environment, biocompatibility. Several research studies have been published in the
excellent combination of strength and Young's modulus appropriate to last decade on the effect of different dopants on the bioactivity of TiO2,
the application, high fatigue and ductility, and excellent biocompat- including Al [9], Mg [10], P [11], N [12], Ca [8,13], Mn [14], Sr [15],
ibility is essential [5,6]. Use of titanium and related alloys, one of the Fe [16,17], Zn [18,19], Nb [20], and Eu [21]. F doping enhances the
most promising engineering classes of materials, is growing rapidly in antibacterial properties of TiO2 [22–24]. However, studies on the


Corresponding author.
E-mail addresses: broy78945@gmail.com, banasri.roy@pilani.bits-pilani.ac.in (B. Roy).

https://doi.org/10.1016/j.jpcs.2019.06.011
Received 21 August 2018; Received in revised form 11 June 2019; Accepted 12 June 2019
Available online 14 June 2019
0022-3697/ © 2019 Elsevier Ltd. All rights reserved.
V. Singh, et al. Journal of Physics and Chemistry of Solids 134 (2019) 262–272

biocompatibility of the fluoride-modified Ti surface have contradictory


and inconclusive results. Some researchers claim that fluoride-doped
TiO2 favors osseointegration and enhances bone tissue bonding
[25–30]. It is believed that the PO34 ion can replace the F− ion and
make anchoring sites for Ca- and P-related compounds on the implants
to facilitate bonding between bone and the implant [31]. Tiainen et al.
[32] investigated the in vitro biological response of murine osteoblasts
to fluoride-modified rutile TiO2. TiO2 disks were prepared by solid-state
sintering of powder, and the fluoride modification was performed by
immersion of the sample disks in dilute hydrofluoric acid (HF). F in-
corporation led to significant changes in the surface topography. At
lower concentration, F doping did not result in any increased cellular
interaction with the coating, whereas at higher concentration, it was
toxic to the osteoblasts. Thor et al. [33] studied the thrombogenic re-
sponse of hydroxyapatite, machined titanium, TiO2 grit–blasted tita-
nium (TiOB), and fluoride-modified grit-blasted titanium (TiOB-F) in
contact with whole blood, platelet-rich plasma, and platelet-poor
plasma. The highest blood coagulation was observed for TiOB and
TiOB-F, with significantly higher activation of the latter. This was ex-
plained by the increased roughness of the implant surface during the F
modification via HF treatment. Kang et al. [34] reported the anti-
bacterial properties and cytocompatibility of a Cl-modified nanos-
tructured TiO2 film. Chlorine incorporation was observed to increase
both the antibacterial effect and cytocompatibility of the structure.
In addition to the chemistry of the material itself, bulk material and
coating interaction, film roughness, homogeneity of the oxide coverage,
and the presence of anchoring sites on the submicron scale or nanoscale Fig. 1. Simultaneous thermogravimetric (thermogravimetric analysis [TGA]
might influence the osteoblast-implant interaction [35]. Proper balance and differential thermal analysis [DTA]) spectra of undoped, Cl-doped, and F-
of the wettability (hydrophobicity and hydrophilicity, polar and dis- doped TiO2.
persive components of surface tension, etc.) of the coating is important
for cell growth and high biocompatibility. Cell adhesion and pro-
decomposition behavior of the dried gels. A 60H-DTA-TG (Shimadzu
liferation can decrease as the material becomes very hydrophilic or
Analytical (India) Pvt. Ltd.) instrument was used for thermogravimetric
hydrophobic.
scanning of the powders in the range from room temperature to 800 °C,
In a previous article we reported on the effect of the dopant type and
with a heating rate of 10 °C/min and under an air atmosphere flowing
concentration on TiO2 powder microstructure via Rietveld analysis
at 25 mL/min. The TGA-DTA results helped us to decide the heat
[36]. The aim of this article is to study the effect of the surface chem-
treatment conditions (time, temperature) for the samples.
istry of Cl- and F-doped TiO2 powder on biocompatibility. X-ray dif-
fraction (XRD) data were used to analyze the phase content and particle
size, and transmission electron microscopy (TEM) images were used to
2.2. Characterization
examine the morphology and particle size distribution. The surface
chemistry and composition were analyzed by X-ray photoelectron
The XRD (Rigaku Minilflex X-ray diffractrometer, Cu Kα radiation,
spectroscopy (XPS). Preliminary hemolysis and cytotoxicity testing
λ = 1.54 Å) spectra of the powders were studied in the 2θ range from
were performed to assess the biocompatibility of the powders. This
10° to 80° at a scanning rate of 0.03°/min. The 2θ versus intensity data
understanding of the materials (interface effect free) gained can be
were analyzed by Rietveld's full spectrum fitting technique based on
extrapolated to realize the effect of Cl or F modification on bio-
both structural and microstructural refinement to understand the effect
compatibility of the TiO2 coatings.
of doping on phase transformation, lattice strain, particle size, etc. The
FullProf software program interfaced with Match! version 3 was used
2. Materials and methods for this purpose. The pseudo-Voigt analytical function was chosen for
fitting the observed XRD scan, because it independently and in-
2.1. Sample preparation dividually considers both the particle size and strain broadening aspects
of the experimental profile [37–39]. The Debye-Scherer formula and
The TiO2 powders were synthesized by a sol-gel route as discussed the full width at half maximum (FWHM) of the XRD 100% peaks were
by Rao et al. [36]. Titanium tetraisoproxide (Ti(C3H5O)4; > 98%, Avra) used to calculate the particle size of the phases. The microstructural
was mixed slowly drop by drop with ethanol (> 99.5%, Changshu strain generated from vacancies, interstitials defects, layer faults, dis-
Yangyuan) in a beaker using a magnetic stirrer at normal ambient locations, etc, was calculated from the slope of the Williamson-Hall plot
temperature. The transparent solution obtained was preserved as the (FWHM × cos θ versus sin θ). The microstructure and surface mor-
control undoped sol. A premeasured amount of hydrochloric acid (HCl; phology of the powders were examined under a transmission electron
37%, Ranchem) or HF (48%, Fischer) was added slowly to this trans- microscope (JEOL 2010; LaB6 filament at 200 kV). The XPS data were
parent sol under continuous stirring to dope the material with Cl or F, collected by a Kratos Axis Ultra DLD X-ray photoelectron spectrometer
respectively. The final clear doped sol had a volume ratio of 1:5:0.05 (Ti using the monochromatic Kα beam of an Al X-ray tube as a radiation
(C3H5O)4 to ethanol to HCl/HF, 10 at% doping). One or two drops of source operating at 13 kV AC voltage and 9 mA emission current and
water were added to the sols to increase the rate of gel formation, and with a best resolution of 0.50 eV. The chamber pressure was maintained
both sols converted to a transparent and colorless gel in 1 day to 1 at approximately 10−9 Torr. The aliphatic carbon peak at 284.8 eV was
week. The gels were dried in a hot air oven at 60 °C for 12 h, and used as a reference for all spectra.
thermogravimetric analysis (TGA) and differential thermal analysis
(DTA) of the powders were performed to understand the thermal

263
V. Singh, et al. Journal of Physics and Chemistry of Solids 134 (2019) 262–272

Fig. 3. Example Rietveld refinement performed on the X-ray diffraction results


for the undoped, Cl-doped, and F-doped TiO2 materials obtained after heat
treatment at 800 °C for 2 h. The black and gray bars at the bottom of each panel
correspond to anatase (JCPDS card no. 00-21-1272) and rutile (JCPDS card no.
00-21-1276) phases, respectively.

2.3. Hemolysis

Blood (10 mL) was drawn from a healthy, nonsmoker and non-
drinker adult human into a heparinized tube containing anticoagulant
(EDTA vacuum tube, BD Biosciences, India). The content of the he-
parinized tube was centrifuged (Eppendorf centrifuge 5415 R, relative
centrifugal force 835g) at 3000 rpm and 4 °C for 20 min. Erythrocytes
were segregated and washed twice with sterilized phosphate-buffered
saline (PBS). The separated red blood cells were suspended in PBS to
achieve hematocrit of 5%. Then 100 μL of sterilized TiO2 powder
samples at different concentrations in PBS (1, 2, 5, and 10 mg/mL) was
sonicated and added to PBS and red blood cell suspension in a volume
ratio of 1:4:15, respectively. The red blood cell mixture was incubated
at 37 °C for 2 h. The postincubation cell mixture was centrifuged for
5 min at 3000 rpm and 4 °C. The absorbance of the supernatant was
measured with a spectrophotometer (Thermo Fisher Scientific,
Multiskan GO). The formula used to calculate the hemolysis ratio (HR)
as a percentage was as follows:

HR (%) = 100 × (Xs − Xn)/(Xp − Xn)

where Xn is the absorbance of the negative control (a mixture of 0.1 mL


erythrocyte suspension and 5 mL PBS), Xs is the absorbance of the
sample supernatant, and Xp is the absorbance of the positive control (a
mixture of 0.1 mL erythrocyte suspension and 5 mL deionized water)

2.4. Cytotoxicity toward HeLa cells measured by the MTT assay

The 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide


(MTT) assay was performed in a 12-well culture plate. HeLa adherent
cells obtained from the National Centre for Cell Science (Pune, India)
were seeded (1.5 × 106 cells per milliliter) in each well with 600 μL
RPMI 1640 complete medium (Life Technologies, India). Then 100 μL
Fig. 2. X-ray diffraction scans of (a) undoped, (b) Cl-doped, and (c) F-doped
TiO2 powders calcined at different temperature and for different times. The
of sterilized TiO2 powder sample suspended in RPMI 1640 complete
effect of the dopants on slowing the phase transformation is clear. medium (5 mg/mL) was added to each well. The sample treated cells
were then incubated in a CO2 incubator for 24 h. The samples were
washed twice with PBS to remove the supernatant in the wells. Then

264
V. Singh, et al. Journal of Physics and Chemistry of Solids 134 (2019) 262–272

20 μL MTT solution (4 mg/mL; Sigma Life Science, USA) and 600 μL

ε (x 103)
RPMI 1640 complete medium were added to each well, and the re-

13.3
sulting mixture was incubated for 4 h in the dark as MTT is light sen-

8.6
1.9
-
-
-
-
-
sitive. The MTT solution was then removed, leaving behind formazan
D (nm) crystals on the living cells. Then 100 μL dimethyl sulfoxide (Sigma Life
Science, India) was added to each well and mixed gently to dissolve all

61
2
9
-
-
-
-
-
formazan crystals. The 12-well plates were read for absorbance at
570 nm with a spectrophotometer (Thermo Fisher Scientific, Multiskan

100
16
%
R

3
-
-
-
-
-
GO). The formula used to calculate cell viability was as follows:
Rietveld refinement results. Phase content (%), particle size (D), and rms strain (ε) present in lattice for undoped, Cl- and F- doped samples heat treated for different time and temperature.

ε(x 10 )
3

Cell viability (%) = [100 × (Asample − Ablank)/(Acontrol − Ablank)]


25.8
23.8
22.5
19.0
where Asample is the absorbance of cells containing the sample, Acontrol is
-
-
-

-
the absorbance of cells (1.5 × 106/mL) suspended in 20 μL MTT solu-
D (nm)

tion (4 mg/mL) and 600 μL RPMI 1640 complete medium, and Ablank is
21
32
36
43

the absorbance of 20 μL MTT solution (4 mg/mL) and 600 μL RPMI


-
-
-

1640 complete medium [40].


An

20
70
97
84
%

-
-
-

-
F-Doped

100
100
100
80
30
%
A

2.5. Cytotoxicity toward peripheral mononuclear blood cells


-
-
-
ε (x 10 )
3

Blood was taken from a healthy donor and collected in a hepar-


10.5
7.1
3.6
2.2

inized tube containing anticoagulant. Sterile PBS was used to dilute the
-
-

-
-

blood in a ratio of 2:1 (PBS to blood). Layer separation of blood com-


D(nm)

ponents was accomplished by use of Histopaque (Histopaque-1077,


Sigma-Aldrich, India). The PBS-diluted blood was layered on
17
65
78
6
-
-

-
-

Histopaque in a 3:1 ratio. The solution containing Histopaque and di-


100

luted blood was centrifuged at 1300 rpm for 30 min. Among the various
19
78
%
R

2
-
-

-
-

layers, lymphocytes (between PBS and Histopaque) were separated


with a syringe and suspended in 10 mL PBS. Diluted lymphocytes were
ε (x 10 )
3

centrifuged two or three times to separate out any other blood com-
28.7
27.8
18.5
19.6

ponents. The clear lymphocytes were suspended in 10 mL fresh PBS and


-

-
-
-

incubated in RPMI 1640 complete medium. Cytotoxicity of the sample


D (nm)

was analyzed in a 12-well plate. The lymphocytes and 100 μL of ster-


22
26
47
42

ilized TiO2 powder sample suspended in RPMI 1640 complete medium


-

-
-
-

were incubated for 24 h in a CO2 incubator at 37 °C. Then 0.4% trypan


An

75
98
81
22
%
Cl-Doped

blue solution (Sigma-Aldrich, India) was used for staining, and cell
-

-
-
-

counting was performed with a hemocytometer. The well containing


100
25
%

only cells and RPMI 1640 complete medium was used as a control. The
A

-
-
-
-
-
-

control well contained 1.1 × 106 cells per milliliter.


ε (x 10 )
3

6.6
3.6
2.7
-
-

-
-
-

3. Results
D (nm)

3.1. TGA and DTA


22
73
98
-
-

-
-
-

‘A’ stands for amorphous; ‘An’ stands for anatase; ‘R’ stands for rutile.

The thermal decomposition and phase transformation behavior


100
23
85
%
R

appear to be different for the three samples (Fig. 1). The endothermic
-
-

-
-

peaks at approximately 90–110 °C denote loss of water and organics


ε (x 10 )
3

loosely absorbed at the surface.


25.5
11.6
19.7

The consequent weight loss was 5 and 9 wt% for the undoped and
-

-
-
-
-

doped samples, respectively. The exothermic peaks (shoulders) be-


D (nm)

tween 200 and 290 °C indicate loss of approximately 15–20 wt% due to
burning of leftover hydrocarbons and other solvents from the material's
29
57
43
-

-
-
-
-

structure. For doped samples, exothermic weight losses at approxi-


100

mately 340 and 395 °C probably correspond to the removal of Cl and F,


An

77
15
%
Undoped

-
-
-
-

respectively, from the matrix, as indicated by XPS. For the undoped


sample, the sharp exothermic peak at approximately 390 °C, accom-
100
%
A

-
-
-
-
-
-
-

panied by zero/minimal weight loss, implies an amorphous to anatase


phase transformation. No weight loss was observed for any samples
Sample treatment (°C/hr)

above 400 °C. The small exothermic peak at approximately 595 °C


suggests the formation of the rutile phase from anatase for the undoped
sample. For the doped samples, no DTA peak for phase transformation
was detected within the range of thermal scanning (up to 800 °C). Heat
treatment at different temperatures and for different times was applied
800/10
800/14
800/20
Table 1

300/2
400/2
600/2
800/2
800/6

for the powders and to characterize them.

265
V. Singh, et al. Journal of Physics and Chemistry of Solids 134 (2019) 262–272

Fig. 4. Transmission electron microscopy images of the (a) undoped, (b) Cl-doped, and (c) F-doped TiO2 powders heat treated under different conditions to obtain (I)
100% anatase (A), (II) anatase plus rutile (A + R), and (III) 100% rutile (R) phases.

3.2. X-ray diffraction 3.3. Transmission electron microscopy

Amorphous, anatase (JCPDS card no. 00-21-1272), and rutile (JCPDS The bright-field TEM images of undoped (Fig. 4a), Cl-doped
card no. 00-21-1276) TiO2 phases were observed to be the sole content of (Fig. 4b), and F-doped (Fig. 4c) powders calcined at various tempera-
the samples (Fig. 2). No other Cl- or F-related secondary phases were de- tures and for various times to obtain 100% anatase, anatase and rutile
tected. For heat treatment at 300 °C for 2 h, the flat nature of the spectrum mixed phase, and 100% rutile phase show that the particle size of the
and no peak evident for the undoped TiO2 (Fig. 2a) show that the powder is TiO2 powders is suppressed by the dopant, and F affects it more than Cl,
amorphous. For heat treatment at 400 °C for 2 h, the powder has the na- which is in accordance with the XRD and TGA-DTA results. High
nocrystalline (FWHM of the anatase 100% peak (101) is approximately crystallinity of the powders is self-evident from the images. F-doped
0.72°) anatase phase. For the powder calcined at 800 °C for 0.5 h, the strong powders reveal a uniform particle size distribution, whereas for un-
peak at approximately 27.5° indicates the presence of a substantial amount doped and Cl-doped powders, a heterogeneous nature of the particle
of the rutile phase. For treatment at 800 °C for 2.5 h, the powder was almost size distribution is observed. The particle size of the Cl-doped sample
totally converted to the rutile phase. The Cl-doped TiO2 (Fig. 2b) sample was 5–40 nm, 10–50 nm, and 40–200 nm for 100% anatase powder
treated at 600 °C for 1 h was only anatase, but for treatment at 800 °C for (Fig. 4b, I), anatase and rutile mixed phase powder (Fig. 4b, II), and
1 h, a small amount of the rutile phase is present, whereas for the sample 100% rutile powder (Fig. 4b, III), respectively. For F-doped powders the
treated at 800 °C for 4 h, comparable amounts of rutile and anatase phases particle size was 2–15 nm, 5–40 nm, and 25–100 nm for 100% anatase
are present. In the case of the F-doped powder (Fig. 2c), amorphous to powder (Fig. 4c, I), anatase and rutile mixed phase powder (Fig. 4c, II),
nanocrystalline anatase (FWHM of the anatase 100% peak (101) is ap- and 100% rutile powder (Fig. 4c, III), respectively. The particle sizes
proximately 1.7°) conversion occurred for treatment at 800 °C for 2 h. Rutile calculated from the TEM images match well with those obtained by
transformation from anatase was detected for treatment at 800 °C for 9 h. XRD.
Clearly, Cl or F incorporation in the lattice reduces the kinetics of the
amorphous to anatase and anatase to rutile TiO2 conversion, and is con-
3.4. X-ray photoelectron spectroscopy
sistent with the TGA-DTA results [36]. Fig. 3 shows an example of Rietveld
refinement performed on the XRD data for the undoped, Cl-doped, and F-
The wide-scan XPS spectra (Figs. 5a and 6a) show the presence of Ti
doped powders treated at 800 °C for 2 h. Table 1 summarizes the Rietveld
(2p, 450–470 eV), O (1s, 527–536 eV), and C (1s, 282–293 eV) for both
refinement results: phase content, particle size, and root-mean-square (rms)
of the doped samples. The C peak decreases in intensity with increasing
strain in the lattice for the samples treated at various temperatures and for
treatment temperature and time. Furthermore, the peaks at approxi-
various times. For the anatase phase, doping enhances the strain, which can
mately 192–204 eV and 680–690 eV indicate the presence of Cl and F,
be attributed to the lattice mismatch that occurred because of substitution of
respectively, for the corresponding doped samples. The high-resolution
Cl− or F−for O2- in TiO2. The ionic radius of
Ti 2p spectra of the doped TiO2 powders (Figs. 5b and 6b) show the
Cl− (r Cl = 1.81 Å, coordination number 6) and F−
presence of two peaks, at 459 eV (usually common for TiO2) and
(r F = 1.19 Å, coordination number 6) is 28% larger and 15% smaller,
respectively, than that of O2- (r O2 = 1.41 Å, coordination number 6 ). 464.7 eV, which indicates that most of the Ti is in the 4 + oxidation
However, in doped rutile phases, strain was calculated to be smaller than in state. For F-doped samples calcined at 400 °C (Fig. 6b, II) and 600 °C
the undoped rutile phase. (Fig. 6b, III) for 2 h, the appearance of a peak at approximately
457.3 eV reveals the presence of Ti3+. With further heat treatment at
higher temperature (800 °C for 2 h), this peak disappeared. For Cl-

266
V. Singh, et al. Journal of Physics and Chemistry of Solids 134 (2019) 262–272

change, and only two peaks were seen for all the samples. The larger
one, at 284.6 eV, implies that most of the C is in the hydrocarbon form
(C–C and/or C–H bond), and the other peak, at 288.2 eV, indicates the
presence of some oxygenated C in O–C=O form.
The Cl (2p3/2) peak at 198.4 eV indicates that Cl is incorporated into
the TiO2 structure as an anion by creating Ti–Cl bonds (Fig. 5e), which
could possibly be formed by substitution for the oxygen in the TiO2
lattice. However, as Cl is observed to disappear from the structure with
high-temperature treatment and because the ionic radius of Cl−
(1.81 Å) is approximately 28% larger than that of O2- (1.41 Å), it can be
concluded that the Cl− ions are chemically bonded with the surface of
TiO2, and are not in the bulk [41–43]. A similar trend was observed for
F-doped samples (Fig. 6e). For the samples calcined at 100 and 400 °C,
the single peak at approximately 685.4 eV indicates that the F− anion is
chemically bonded with Ti as a Ti–F structure [44,45]. The intensity of
the peak obtained for the powder calcined at 400 °C for 2 h is lower
than that of the peak for the powder calcined at 100 °C for 2 h. This
show that both Cl and F (r F = 1.19 Å) dopants decrease for the TiO2
powder calcined at the higher temperature, although in the samples
treated under the same conditions (100 °C for 2 h or 400 °C for 2 h) F
appears to be present in higher amounts than Cl, which is because the
Ti–F bond (bond dissociation energy at 25 °C of 569 ± 33 kJ/mol) is
much stronger than the Ti–Cl bond (bond dissociation energy at 25 °C of
405.4 ± 10 kJ/mol) [46]. Energy-dispersive X-ray spectroscopy of F-
and Cl-doped TiO2 samples treated under different thermal conditions
showed a similar trend of losing dopants with increasing treatment
temperature [36]. Table 2 summarizes the XPS elemental composition
of the samples obtained after different heat treatments.

3.5. Hemolysis

For all samples, in general, HR increases with the powder con-


centration (Fig. 7a). The effects of doping and the treatment conditions
on the hemolytic activity of the TiO2 samples are clear. At lower
treatment temperatures, doped anatase TiO2 powders show higher HR
than the undoped anatase sample. Even at the highest concentration
(10 mg/mL PBS) the undoped sample appears to be nontoxic (HR <
5%), but the doped powders show toxicity (HR > 5%). In the mixed
phase powders, the presence of the thermodynamically stabler rutile is
observed to help reduce HR (Fig. 7a and b). The doped rutile phase
powders show less hemolytic activity than the undoped rutile TiO2. The
F-doped rutile sample exhibits the lowest hemolytic activity at all
concentrations and is nonhemolytic (HR < 2%), even at the highest
Fig. 5. X-ray photoelectron spectroscopy spectra of Cl-doped TiO2 powders
concentration (10 mg/mL PBS). It shows hemoglobin release the same
heat treated at (I) 100 °C for 2 h, (II) 400 °C for 2 h, (III) 600 °C for 2 h, and (IV)
800 °C for 2 h: (a) wide scan, (b) Ti 2p, (c) O 1s, (d) C 1s, and (e) Cl 2p. as that of the negative control and clearly less than that of the Cl-doped
and undoped rutile powders (Fig. 7c).

doped powder (Fig. 5b) or undoped powder (Supplementary Data


3.6. Cytotoxicity
Fig. 1), the presence of Ti3+ is not observed.
The O 1s spectra have three peaks (Figs. 5c and 6c). The peak at
The in vitro cytotoxicity of the undoped and doped TiO2 samples at
approximately 530.0 ± 0.2 eV is typical of the Ti–O bond. At 100 °C
5 mg/mL RPMI 1640 powder concentration toward HeLa cells mea-
this peak is weak and wide, but its intensity increases as the samples are
sured by the MTT assay (Fig. 8a) and toward peripheral mononuclear
calcined at higher temperatures. The peak at approximately
blood cells (PMBCs) (Fig. 9a) shows a similar trend. The doped anatase
532 ± 0.4 eV was attributed to the O atoms from O–H. For the powder
samples appear to be more toxic (HeLa cell survival was 40% and 30%
calcined at 100 °C for 2 h, the peak is strong, but becomes weaker with
for Cl- and F-doped powders, respectively, and PMBC survival was 51%
further heat treatment, but the intensity of the peak is noticeably higher
and 44% for Cl- and F-doped powders, respectively) than the undoped
for the Cl-doped powder than that of the F-doped powder, which might
anatase sample (HeLa cell survival was 60% and PMBC survival was
imply a more hydrophilic nature of the Cl-doped powder. The small
58%). However, for the doped rutile samples (Figs. 8b and 9b) the cell
peak at approximately 529 eV for the F-doped powder calcined at
survival rate was higher (HeLa cell survival was 86% and 97% for Cl-
400 °C for 2 h and at 600 °C for 2 h is probably due to Ti3+.
and F-doped powders, respectively, and PMBC survival was 84% and
Three peaks appear in the C 1s spectra of Cl- and F-doped powders
90% for Cl- and F-doped powders, respectively) than for undoped rutile
calcined at 100 °C for 2 h (Figs. 5d and 6d). The largest one, at
(HeLa cell survival was 83.5% and PMBC survival was 83%).
285.5 ± 0.2 eV, was assigned to C in C–OH and/or COOH. The second
largest peak, at 287.3 eV, was assigned to carbonyl C in C=O, and the
4. Discussion
smallest one, at 289.5 eV, was assigned to carboxylate C in O–C=O.
With further heat treatment at higher temperatures, the peak positions
Cl or F ion can substitute for O2- ion on the surface of the particles,

267
V. Singh, et al. Journal of Physics and Chemistry of Solids 134 (2019) 262–272

Fig. 6. X-ray photoelectron spectroscopy spectra of F-doped TiO2 powders heat treated at (I) 100 °C for 2 h, (II) 400 °C for 2 h, (III) 600 °C for 2 h, and (IV) 800 °C for
2 h: (a) wide scan, (b) Ti 2p, (c) O 1s, (d) C 1s, and (e) F 1s.

create lattice distortion, and create defects, and as a result inhibit the spectrum acquired from the F-doped powders after treatment at 100 °C
phase transformations and at the same time refine the particle size, with for 2 h and at 400 °C for 2 h, F doping substitutes for oxygen and
F being more effective than Cl. As is evident from the Ti 2p XPS transforms some Ti4+ to Ti3+ to conserve charge neutrality, which in

268
V. Singh, et al. Journal of Physics and Chemistry of Solids 134 (2019) 262–272

Table 2
XPS elemental composition of the samples obtained after different heat treatments.
Sample treatment (°C/hr) Undoped Cl-Doped F-Doped

Ti (at%) O (at%) C (at%) Ti (at%) O (at%) C (at%) Cl (at%) Ti (at%) O (at%) C (at%) F (at%)

100/2 17 50 33 18 44 31 7 18 39 32 11
400/2 19 60 21 21 58.7 19 1.3 22 56.5 18 3.5
600/2 22 60 18 23 61 16 - 24 62 14 -
800/2 24 62 14 25 62 13 - 26 63 11 -

Fig. 7. Hemolytic activity of the undoped, Cl-doped,


and F-doped TiO2 powders: (a) the effect of different
powder concentrations and different phases; (b) the
effect of different heat treatment conditions for
5 mg/mL phosphate-buffered saline (PBS) powder
concentration; (c) photographs showing the hemo-
lytic activity for 100% rutile powders (5 mg/mL PBS)
with positive and negative controls. HR, hemolysis
ratio.

turn creates extra oxygen vacancies in the structure: more defect-free and less toxic. On the basis of XRD and high-resolution
TEM findings, many researchers have concluded that F doping induces
1 a smaller grain size, but improves the crystallinity (induces fewer de-
O×O + 2Ti×Ti O2 + 2Ti Ti + VO••.
2 fects) of anatase TiO2 [51–53]. Xu and Zhang [54] observed the same
for Cl-doped anatase TiO2. Anatase TiO2 with a high concentration of Cl
This is quite consistent with the findings of other researchers
doping has higher crystallinity. Relatively fewer studies have been re-
[47–49]. Dozzi et al. [49] studied the time-resolved photoluminescence
ported on the effect of F or Cl (or other halogen) doping on the crys-
spectrum of F-doped anatase TiO2 powder prepared by a sol-gel method
tallinity of rutile TiO2.
using HF as a dopant source (as in the present study). The photo-
Rietveld analysis shows that the rms strain in the undoped (after
luminescence signal suggested the accumulation of heavy defects on the
treatment at 400 °C for 2 h), Cl-doped (after treatment at 600 °C for 3 h),
material surface, which could be due to the creation of Ti3+ centers in
and F-doped (after treatment at 800 C for 9 h) anatase powders is 26,
the lattice. Li et al. [50] used photoluminescence and ESR spectroscopy
29, and 35, respectively. Under those heat treatment conditions, doped
to study the effect of F doping in anatase TiO2, and showed that oxygen
samples do not contain Cl and F in the lattice but still have higher strain
vacancies with trivalent Ti3+ ions could be created in the TiO2 lattice
and slower phase transformation kinetics, indicating the presence of
due to the F doping. The NH3 temperature-programmed desorption
defects in the lattice. With further heat treatment, the particle size in-
characterization showed that F doping enhances the generation of
creases, and as a result, strain in the structure decreases, but in a dif-
acidic sites on the anatase TiO2 surface. It also supports the idea that F
ferent manner. Different powders convert to 100% rutile phase under
in the TiO2 structure enhances the creation of O2 radicals. The presence
different treatment conditions (Table 1). The rms strain was calculated
of active radicals and surface defects could be the reason for the higher
to be 2.7, 2.2, and 1.9 for the undoped, Cl-doped, and F-doped rutile
surface toxicity of F-doped anatase samples and explain the high HR
powders, respectively. It is possible that the higher-energy defects
and cytotoxicity of these powders. There is no F left in the samples
present in the doped crystals act as new nucleation sites and a high
calcined at 600 °C or above and that can explain why no more Ti3+ is
treatment temperature provides the ions with greater mobility to re-
observed in these samples, which presumably indicates that as the
arrange themselves. This leads to the formation of a new rutile phase
amount of the rutile phase increases the samples become more and

269
V. Singh, et al. Journal of Physics and Chemistry of Solids 134 (2019) 262–272

Fig. 8. Cytotoxicity toward HeLa cells


measured by the 3-(4,5-dimethylthiazol-2-
yl)-2,5-diphenyl tetrazolium bromide
(MTT) assay for the undoped, Cl-doped, and
F-doped TiO2 samples at 5 mg/mL RPMI
1640 powder concentration: (a) the effect of
different heat treatment conditions and (b)
photographs of the formazan crystals on the
living cells for the 100% rutile powders.

Fig. 9. Cytotoxicity toward peripheral mononuclear blood cells for the undoped, Cl-doped, and F-doped TiO2 samples at 5 mg/mL RPMI 1640 powder concentration:
(a) the effect of different heat treatment conditions and (b) photographs of the well plates after incubation with different powders as mentioned.

with less strain and greater thermodynamic stability. The slower ki- doped anatase powders compared with the undoped anatase powder.
netics of the phase transformation for the doped powders may also have Reduced strain implies lower stress and lower reactivity of the lattice.
some effect in lowering the strain. Strain field is related to the presence This defect-free, less toxic, and less reactive surface could be the reason
of point defects in materials and controls the chemical properties for the enhanced biocompatibility of F-and Cl-doped rutile powders
[55,56]. Higher strain in the structure indicates the presence of larger compared with the undoped rutile powder [36].
stress, which leads to higher Brønsted acidity of rutile TiO2 as indicated
by Wang et al. [57]. This could explain the greater bioactivity of the

270
V. Singh, et al. Journal of Physics and Chemistry of Solids 134 (2019) 262–272

5. Conclusion 85139–85152.
[17] Y. Yan, J. Sun, Y. Han, D. Li, K. Cui, Microstructure and bioactivity of Ca, P and Sr
doped TiO2 coating formed on porous titanium by micro-arc oxidation, Surf. Coat.
Cl- or F-doping significantly influences the phase transition kinetics, Technol. 205 (2010) 1702–1713.
particle size, and surface chemistry of TiO2 powder. Incorporation of F [18] H. Hu, W. Zhang, Y. Qiao, X. Jiang, X. Liu, C. Ding, Antibacterial activity and in-
or Cl in the lattice at low temperature creates surface defects and acidic creased bone marrow stem cell functions of Zn-incorporated TiO2 coatings on ti-
tanium, Acta Biomater. 8 (2012) 904–915.
sites, which enhance the negative bioactivity of the anatase phase in- [19] K. Huo, X. Zhang, H. Wang, L. Zhao, X. Liu, P.K. Chu, Osteogenic activity and an-
itially. After treatment at a higher temperature, the dopant ions escape tibacterial effects on titanium surfaces modified with Zn-incorporated nanotube
the lattice and the doped rutile becomes less strained and shows rela- arrays, Biomaterials 34 (2013) 3467–3478.
[20] D. Ding, C. Ning, L. Huang, F. Jin, Y. Hao, S. Bai, Y. Li, M. Li, D. Mao, Anodic
tively better biocompatibility compared with the undoped TiO2 fabrication and bioactivity of Nb-doped TiO2 nanotubes, Nanotechnology 20 (2009)
powder. 305103.
[21] K. Cheng, Y. Zhu, W. Weng, J. Lin, H. Wang, Biocompatible Eu-doped TiO2 nanodot
film with in situ protein adsorption characterization property, Thin Solid Films 584
Acknowledgments
(2015) 9–12.
[22] C.P. Peremarch, R.P. Tanoira, M.A. Arenas, E. Matykina, A. Conde, J.J. De
The authors acknowledge the Micro and Nano Characterization Damborenea, E.G. Barrena, J. Esteban, Bacterial adherence to anodized titanium
Facility, Centre for Nano Science and Engineering, IISc Bangalore, for alloy, J. Phys. Conf. Ser. 252 (2010) 012011.
[23] C. Pérez‐Jorge, A. Conde, M.A. Arenas, R. Pérez‐Tanoira, E. Matykina, J.J. de
XPS characterization, the Physics Department, BITS Pilani, Pilani Damborenea, E. Gómez‐Barrena, J. Esteban, In vitro assessment of Staphylococcus
Campus, for XRD analysis, and BITS Pilani, Pilani Campus, for financial epidermidis and Staphylococcus aureus adhesion on TiO2 nanotubes on Ti–6Al–4V
assistance under an OPERA award and RI grant. alloy, J. Biomed. Mater. Res. A 100 (2012) 1696–1705.
[24] M.A. Arenas, C. Pérez-Jorge, A. Conde, E. Matykina, J.M. Hernández-López,
R. Pérez-Tanoira, J.J. de Damborenea, E. Gómez-Barrena, J. Esteba, Doped TiO2
Appendix A. Supplementary data anodic layers of enhanced antibacterial properties, Colloids Surf., B 105 (2013)
106–112.
[25] M. Monjo, S.F. Lamolle, S.P. Lyngstadaas, H.J. Rønold, J.E. Ellingsen, In vivo ex-
Supplementary data to this article can be found online at https:// pression of osteogenic markers and bone mineral density at the surface of fluoride-
doi.org/10.1016/j.jpcs.2019.06.011. modified titanium implants, Biomaterials 29 (2008) 3771–3780.
[26] S.F. Lamolle, M. Monjo, S.P. Lyngstadaas, J.E. Ellingsen, H.J. Haugen, Titanium
implant surface modification by cathodic reduction in hydrofluoric acid: surface
Conflicts of interest characterization and in vivo performance, J. Biomed. Mater. Res. A 88 (2009)
581–588.
The authors declare that they have no conflict of interest. [27] J.E. Ellingsen, Pre-treatment of titanium implants with fluoride improves their re-
tention in bone, J. Mater. Sci. Mater. Med. 6 (1995) 749–753.
[28] C.G. Bellows, J.N. Heersche, J.E. Aubin, The effects of fluoride on osteoblast pro-
References genitors in vitro, J. Bone Miner. Res. 5 (1990) S101–S105.
[29] J.E. Wergedal, K.H. Lau, D.J. Baylink, Fluoride and bovine bone extract influence
[1] D. Landolt, Electrochemical and materials aspects of tribocorrosion systems, J. cell proliferation and phosphatase activities in human bone cell cultures, Clin.
Phys. D Appl. Phys. 39 (2006) 3121–3127. Orthop. Relat. Res. 233 (1988) 274–282.
[2] P. Ponthiaux, F. Wenger, D. Drees, J.P. Celis, Electrochemical techniques for [30] L.F. Cooper, Y. Zhou, J. Takebe, J. Guo, A. Abron, A. Holmén, J.E. Ellingsen,
studying tribocorrosion processes, Wear 256 (2004) 459–468. Fluoride modification effects on osteoblast behavior and bone formation at TiO2
[3] Y. Yan, A. Neville, D. Dowson, Biotribocorrosion—an appraisal of the time de- grit-blasted cp titanium endosseous implants, Biomaterials 27 (2006) 926–936.
pendence of wear and corrosion interactions: I. The role of corrosion, J. Phys. D [31] Z.M. Isa, G.B. Schneider, R. Zaharias, D. Seabold, C.M. Stanford, Effects of fluoride-
Appl. Phys. 39 (2006) 3200. modified titanium surfaces on osteoblast proliferation and gene expression, Int. J.
[4] J.L. Gilbert, S. Sivan, Y. Liu, S.B. Kocagöz, C.M. Arnholt, S.M. Kurtz, Direct in vivo Oral Maxillofac. Implant. 21 (2006) 203–211.
inflammatory cell‐induced corrosion of CoCrMo alloy orthopedic implant surfaces, [32] H. Tiainen, M. Monjo, J. Knychala, O. Nilsen, S.P. Lyngstadaas, J.E. Ellingsen,
J. Biomed. Mater. Res. A 103 (2015) 211–223. H.J. Haugen, The effect of fluoride surface modification of ceramic TiO2 on the
[5] S.A. Brown, J.S. Kawalec, A.C. Montague, K. Merritt, J.H. Payer, Effects of material surface properties and biological response of osteoblastic cells in vitro, Biomed.
combination, surface treatment, and environment on fretting corrosion of Ti 6Al 4V, Mater. 6 (2011) 045006.
in: S.A. Brown, J.E. Lemons (Eds.), Medical Applications of Titanium and its Alloys: [33] A. Thor, L. Rasmusson, A. Wennerberg, P. Thomsen, J.M. Hirsch, B. Nilsson,
the Material and Biological Issues, ASTM International, West Conshohocken, 1996, J. Hong, The role of whole blood in thrombin generation in contact with various
pp. 231–239. titanium surfaces, Biomaterials 28 (2007) 966–974.
[6] M.G. Manda, P.P. Psyllaki, D.N. Tsipas, P.T. Koidis, Observations on an in‐vivo [34] M.-K. Kang, S.-K. Moon, K.-M. Kim, K.-N. Kim, Antibacterial effect and cyto-
failure of a titanium dental implant/abutment screw system: a case report, J. compatibility of nano-structured TiO2 film containing Cl, Dent. Mater. J. 30 (2011)
Biomed. Mater. Res. B 89 (2009) 264–273. 790–798.
[7] J.H. Kim, H.S. Lee, J.C. Goak, Y.H. Seo, K.B. Kim, K.S. Park, N.S. Lee, Effect of micro [35] D. Lozano, J.M. Hernández‐López, P. Esbrit, M.A. Arenas, E. Gómez‐Barrena, J. de
and nanoparticle inorganic fillers on the field emission characteristics of photo- Damborenea, J. Esteban, C. Pérez‐Jorge, R. Pérez‐Tanoira, A. Conde, Influence of
sensitive carbon nanotube pastes, Appl. Surf. Sci. 256 (2010) 2636–2642. the nanostructure of F‐doped TiO2 films on osteoblast growth and function, J.
[8] J.W. Park, K.B. Park, J.Y. Suh, Effects of calcium ion incorporation on bone healing Biomed. Mater. Res. A 103 (2015) 1985–1990.
of Ti6Al4V alloy implants in rabbit tibiae, Biomaterials 28 (2007) 3306–3313. [36] A. Rao, V.S. Pundir, A. Tiwari, Y. Padarthi, N.V.M. Rao, S. Aich, B. Roy,
[9] S. Aich, M.K. Mishra, C. Sekhar, D. Satapathy, B. Roy, Synthesis of Al-doped Nano Investigating the effect of dopant type and concentration on TiO2 powder micro-
Ti-O scaffolds using a hydrothermal route on titanium foil for biomedical applica- structure via Rietveld analysis, J. Phys. Chem. Solids 13 (2018) 164–176.
tions, Mater. Lett. 178 (2016) 135–139. [37] H.M. Rietveld, Line profiles of neutron powder-diffraction peaks for structure re-
[10] J.W. Park, Y.J. Kim, J.H. Jang, C.H., An, in vitro biocompatibility of magnesium- finement, Acta Crystallogr. 22 (1967) 151–152.
incorporated submicro-porous titanium oxide surface produced by hydrothermal [38] H.M. Rietveld, A profile refinement method for nuclear and magnetic structures, J.
treatment, Appl. Surf. Sci. 257 (2010) 925–931. Appl. Crystallogr. 2 (1969) 65–71.
[11] P. Yang, Y.X. Leng, A.S. Zhao, H.F. Zhou, L.X. Xu, S. Hong, N. Huang, [39] R.A. Young, Introduction to the Rietveld method, in: R.A. Young (Ed.), The Rietveld
Bloodcompatibility improvement of titanium oxide film modified by phosphorus Method, Oxford University Press/IUCr, Oxford, 1993, pp. 1–38.
ion implantation, Nucl. Instrum. Methods B 242 (2006) 15–17. [40] R.I. Freshney, Culture of Animal Cells: A Manual of Basic Technique and Specialized
[12] M. Kawashita, N. Endo, T. Watanabe, T. Miyazaki, M. Furuya, K. Yokota, Y. Abiko, Applications, sixth ed., Wiley-Blackwell, Oxford, 2010.
H. Kanetaka, N. Takahashi, Formation of bioactive N-doped TiO2 on Ti with visible [41] Y. Cao, Y. Yu, P. Zhang, L. Zhang, T. He, Y. Cao, An enhanced visible-light pho-
light-induced antibacterial activity using NaOH, hot water, and subsequent am- tocatalytic activity of TiO2 by nitrogen and nickel–chlorine modification, Sep. Purif.
monia atmospheric heat treatment, Colloids Surf., B 145 (2016) 285–290. Technol. 104 (2013) 256–262.
[13] B. Burnat, J. Robak, D. Batory, A. Leniart, I. Piwoński, S. Skrzypek, M. Brycht, [42] X.K. Wang, C. Wang, W.Q. Jiang, W.L. Guo, J.G. Wang, Sonochemical synthesis and
Surface characterization, corrosion properties and bioactivity of Ca-doped TiO2 characterization of Cl-doped TiO2 and its application in the photodegradation of
coatings for biomedical applications, Surf. Coat. Technol. 280 (2015) 291–300. phthalate ester under visible light irradiation, Chem. Eng. J. 189 (2012) 288–294.
[14] J.W. Park, Y.J. Kim, J.H. Jang, Surface characteristics and in vitro biocompatibility [43] H. Xu, Z. Zheng, L. Zhang, H. Zhang, F. Deng, Hierarchical chlorine-doped rutile
of a manganese-containing titanium oxide surface, Appl. Surf. Sci. 258 (2015) TiO2 spherical clusters of nanorods: large-scale synthesis and high photocatalytic
977–985. activity, J. Solid State Chem. 181 (2008) 2516–2522.
[15] J.W. Park, Increased bone apposition on a titanium oxide surface incorporating [44] S. Zhang, Q. Zhong, W. Zhao, Y. Li, Surface characterization studies on F-doped
phosphate and strontium, Clin. Oral Implant. Res. 22 (2011) 230–234. V2O5/TiO2 catalyst for NO reduction with NH3 at low-temperature, Chem. Eng. J.
[16] D. Flak, E. Coy, G. Nowaczyk, L. Yate, S. Jurga, Tuning the photodynamic efficiency 253 (2014) 207–216.
of TiO2 nanotubes against HeLa cancer cells by Fe-doping, RSC Adv. 5 (2015) [45] Y. Zhang, C. Han, M.N. Nadagouda, D.D. Dionysiou, The fabrication of innovative
single crystal N, F-codoped titanium dioxide nanowires with enhanced

271
V. Singh, et al. Journal of Physics and Chemistry of Solids 134 (2019) 262–272

photocatalytic activity for degradation of atrazine, Appl. Catal., B 168 (2015) Langmuir 15 (1999) 5422–5425.
550–558. [52] J. Yu, C.Y. Jimmy, M.K. Leung, W. Ho, B. Cheng, X. Zhao, J. Zhao, Effects of acidic
[46] D.R. Lide, Molecular structure and spectroscopy, in: W.M. Haynes (Ed.), CRC and basic hydrolysis catalysts on the photocatalytic activity and microstructures of
Handbook of Chemistry and Physics, CRC Press, Boca Raton, 2010, pp. 50–58. bimodal mesoporous titania, J. Catal. 217 (2003) 69–78.
[47] S. Tosoni, D.F. Hevia, O.G. Diaz, F. Illas, Origin of optical excitations in fluorine- [53] T. Giannakopoulou, N. Todorova, C. Trapalis, T. Vaimakis, Effect of fluorine doping
doped titania from response function theory: relevance to photocatalysis, J. Phys. and SiO2 under-layer on the optical properties of TiO2 thin films, Mater. Lett. 61
Chem. Lett. 3 (2012) 2269–2274. (2007) 4474–4477.
[48] Y. Wang, H. Zhang, P. Liu, T. Sun, Y. Li, H. Yang, X. Yao, H. Zhao, Nature of visible- [54] H. Xu, L. Zhang, Selective nonaqueous synthesis of C−Cl-codoped TiO2 with
light responsive fluorinated titanium dioxides, J. Mater. Chem. A 1 (2012) visible-light photocatalytic activity, J. Phys. Chem. C 114 (2010) 11534–11541.
12948–12953. [55] B.R. Ortiz, H. Peng, A. Lopez, P.A. Parilla, S. Lany, E.S. Toberer, Effect of extended
[49] M.V. Dozzi, C. D'Andrea, B. Ohtani, G. Valentini, E. Selli, Fluorine-doped TiO2 strain fields on point defect phonon scattering in thermoelectric materials, Phys.
materials: photocatalytic activity vs time-resolved photoluminescence, J. Phys. Chem. Chem. Phys. 17 (2015) 19410–19423.
Chem. C 117 (2013) 25586–25595. [56] D.V. Potapenko, Z. Li, J.W. Kysar, R.M. Osgood, Nanoscale strain engineering on the
[50] D. Li, N. Ohashi, S. Hishita, T. Kolodiazhnyi, H. Haneda, Origin of visible-light- surface of a bulk TiO2 crystal, Nano let 14 (2014) 6185–6189.
driven photocatalysis: a comparative study on N/F-doped and N–F-codoped TiO2 [57] Q. Wang, A.R. Oganov, O.D. Feya, Q. Zhu, D. Ma, The unexpectedly rich re-
powders by means of experimental characterizations and theoretical calculations, J. constructions of rutile TiO2(011)-(2 × 1) surface and the driving forces behind
Solid State Chem. 178 (2005) 3293–3302. their formation: an ab initio evolutionary study, Phys. Chem. Chem. Phys. 18
[51] A. Hattori, K. Shimoda, H. Tada, S. Ito, Photoreactivity of sol-gel TiO2 films formed (2016) 19549–19556.
on soda-lime glass substrates: effect of SiO2 underlayer containing fluorine,

272

You might also like