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

Ceramics International 49 (2023) 36265–36275

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

Structural, electronic and optical characteristics of TiO2 and Cu-TiO2 thin


films produced by sol-gel spin coating
K. Albaidani a, **, A. Timoumi a, ***, W. Belhadj a, *, S.N. Alamri b, Saleh A. Ahmed c, d
a
Department of Physics, Faculty of Applied Science, Umm AL-Qura University, Makkah, 24382, Saudi Arabia
b
Department of Physics, Science Faculty, Taiba University, Madinah, Saudi Arabia
c
Department of Chemistry, Faculty of Applied Science, Umm Al-Qura University, 21955, Makkah, Saudi Arabia
d
Department of Chemistry, Faculty of Science, Assiut University, 71516, Assiut, Egypt

A R T I C L E I N F O A B S T R A C T

Handling Editor: Dr P. Vincenzini Sol-gel spin coating was employed to deposit titanium dioxide (TiO2) and Cu-TiO2 composites on glass substrates.
Cu was utilized as a dopant in varying amounts. The structural, morphological, optical, and dielectric charac­
Keywords: teristics of the produced samples were evaluated using a variety of techniques. XRD analysis was used to confirm
Cu-TiO2 the anatase (TiO2) phase’s existence. After adding Cu(NO3)2.3(H2O) to TiO2, it was observed that Cu atoms were
Copper
mostly distributed on the TiO surface, resulting in a decrease in the particle size, as revealed by the TEM images.
Sol-gel spin coating
Raman Spectroscopy and FTIR analysis indicates the expansion of the lattice TiO2 with Cu amount and the
Titanium dioxide
Thin films formation of vibrating Ti-O-Ti bonds respectively. With increasing dopants, spherical nanoparticles begin to form
DFT+U correction and orient themselves to aggregate, as shown in the AFM’s noticeable shift in roughness. The UV–vis spec­
Photovoltaic troscopy revealed a shift of the absorption peak toward visible range that signified an increase in the bandgap
after doping. The theoretical study using quantum espresso ab initio simulation and DFT + U correction were
used to derive the electronic band gap energy values which are consistent with our experimental results.

1. Introduction deposition (ALD) [6], sol-gel methods [7], and others [8,9]. The sol-gel
technique that produces TiO2 thin films is based on the hydrolysis and
Because of their particular physical, chemical, optical, and opto­ polycondensation of titanium alkoxides combined with alcohol and
electronic features, transparent conducting oxide (TCO) composites catalytic chemicals. Ti alkoxides should be utilized in conjunction with
such as TiO2, ZnO, SnO2, and CdO are of significant concern. TiO2 is the their corresponding alcohol. The precursor solution, is a colloidal sus­
most promising among these materials in several fields of study due to pension of Ti surrounded by ligands with sufficient physical-chemical
its photocatalytic efficacy, large refractive index, chemical stability, low characteristics to form a film. The film is generated by a wet gel that
cost, and non-toxicity [1]. Thin films find their applications in the op­ becomes a dry gel following deposition, which can be done via
toelectronics domain, photonics and electronics, as the layer of mate­ dip-coating, spin-coating, or spray-coating techniques. Because the re­
rials exhibits more improved physical characteristics than the bulk [2]. actions occur concurrently throughout the deposition phase, this
The most common polymorphs, brookite, rutile, and anatase phases, mechanism is very complicated [10]. In recent years, much work has
exhibit direct optical gaps of roughly 3.37 eV, 3.4 eV, and 3.2 eV, gone into microscopic surface engineering of the anatase crystal to
respectively, according to the investigation of the optical transition in better understand the connection between surface and photocatalytic
high-fraction titania (TiO2) thin films [3]. The anatase phase is more activity. The energy needs for electron transfer and light absorption are
active in photocatalysis [4]. Numerous branches of science and tech­ significantly influenced by the placement of the band gap as well as the
nology use thin film deposition in a variety of ways. To achieve this band edge magnitude throughout photocatalytic reactions. According to
diversity of applications, several preparation methods have been research studies on the bandgap and band edge position in TiO2, a sig­
designed including chemical vapor deposition (CVD) [5], atomic layer nificant limiting factor for TiO2 photocatalytic activity is the rate of

* Corresponding author.
** Corresponding author.
*** Corresponding author.
E-mail addresses: s44285041@st.uqu.edu.sa (K. Albaidani), aotemoume@uqu.edu.sa (A. Timoumi), wbbelhadj@uqu.edu.sa (W. Belhadj).

https://doi.org/10.1016/j.ceramint.2023.08.309
Received 19 November 2022; Received in revised form 24 August 2023; Accepted 28 August 2023
Available online 29 August 2023
0272-8842/© 2023 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
K. Albaidani et al. Ceramics International 49 (2023) 36265–36275

recombination of photogenerated electron-hole pairs [11]. Electronics, on square shaped substrates placed in an oven at 400 ◦ C for about 1 h.
optics, environment, protection, and medicine are just a few of the fields The thickness of the thin films obtained lies between (700–1000) nm.
where TiO2 thin films are used. In thin films, only the tetragonal phases
anatase and rutile TiO2 have been found. TiO2 cause photoreactivity by 2.2. Characterization technique
changing the interfacial electron transfer by doping [12,13]. Doping
TiO2 with metallic ions causes changes in its photoresponse property Obtained samples were characterized by X-Ray diffractometer
[14]. The physical properties of TiO2 thin films could be strongly (RIGAKU ULTIMA-IV). We have used an X-Ray CuKα-source radiating at
modified by doping with Fe, Ag, Zn and Cu [15–18]. TiO2 doped with λ = 1.5406 Å with scanning angle ranging from 10 to 80◦ . The Raman
Nb, Fe, and Zn [19–21] is one of the most important metal in lowering spectra were measured using a Raman Spectroscopy (Thermo Fisher
the bandgap energy of TiO2 and making it a visible light active material. Scientific, model: DXR Raman Microscope) equipped with an argon laser
Few works on the applications of Cu doped TiO2 have been reported in (532 nm). The FTIR measurements were made by FTIR technique
the areas of solar energy [22]. Copper (Cu) is a useful dopant for TiO2 (model: FTIR - IRAffinity1S, Shimadzu). JEOL JSM-7600F. The used
because its atomic size is similar to that of titanium. As a result, it dis­ SEM is equipped with EDS (LEO 1530) FEG–SEM–EDS, and it was uti­
solves quickly in it. Cu plays a part in TiO2 by capturing and transferring lized to examine the surface shape and determine its elemental ratio. For
photo-excited electrons on the material’s surface. Cu should be posi­ the 2D and 3D AFM morphology, we have used a Veeco CP-II AFM in
tioned in Ti locations after being doped into TiO2 and then heated to a contact mode with scan rate (1 Hz). The XPS was performed using a JPS-
high temperature (500–600 ◦ C). Additionally, it has been noted that 9030 Photoelectron Spectrometer with an Al-Kα X-ray source which
adding Cu to the TiO2 matrix may have the benefit of reducing charge provide photons with 1487 eV. All binding energies have been bench­
carrier recombination [23–25]. marked to Oxygen (1s) at 530.7 eV or to Carbon (1s) at 284.6 eV. With a
In the current work, various physical properties of TiO2 and Cu-TiO2 PHILIPS CM-200 electron microscope (Thermo scientific, Themis 300
thin films produced by the sol-gel spin coating method will be investi­ G3) and a field emission electron microscope analysis, transmission
gated and carefully analyzed. The effect of the deposited films with electron microscopy (TEM) with a 200 kV accelerating voltage was used
different amounts of copper on their structural, morphological, optical to examine the morphology of our samples. The optical spectra were
and electronic properties were studied and analyzed in detail by both measured at normal incidence and room temperature for the range 200-
experimental and theoretical methods. 3500 nm at a 0.1 nm resolution using a spectrophotometer type (Shi­
madzu Solid Spec-3700 ultraviolet–visible with an integrating sphere).
2. Materials and experimental study The thickness was determined using a stylus displacement of a Veeco­
Dektak 150 profilometer.
The chemicals employed in this work are all of analytical quality and
weren’t further purified before usage. Titanium (IV) Isopropoxile (TIP) 3. Results and discussion
(Ti4O28H12C) (99.999%), Copper (II) nitrate hydrate (Cu(NO3)2.3(H2O))
(99.999%), Ethanol (≥99.8%) and acetic acid (≥99.8%) were supplied 3.1. XRD analysis
by Sigma Aldrich.
Firstly, the type of TiO2 crystallinity is identified using XRD analysis.
2.1. Synthesis of the samples The Crystallite size was calculated by the formula of Eq. (1) [2,28,
30–32] and the results are displayed in Table 1.
The preparation procedure of TiO2 samples (see steps in Fig. 1) was kλ
previously explained and detailed by Timoumi et al. [27]. Thin films D= (1)
β cos θ
studied in this work were produced employing sol-gel spin coating
methodology. For doping; various quantities (0.1, 0.2 and 0.3g) of (Cu here D represents crystallite size, k is known as Scherer’s parameter (k =
(NO3)2.3(H2O)) were added to the TiO2 solution. The first sample was 0.94) and β is the full width at half maximum (FWHM). Table 1 reveals
prepared without copper. Subsequently 2 ml of liquid were placed on a that the existence of copper slightly reduced the crystallite size of TiO2.
glass substrate (1 cm × 2 cm), then the rotation was used is 6000 rpm for The smaller size of TiO2 crystallites could be attributed to the suppres­
20 s. The glass substrate was widely used because it was readily avail­ sion of titania condensation and crystallization in the Cu-doped solution
able, inexpensive, and easy to clean. Finally, the samples were deposited [29]. Refinements have been made using a tetragonal structure with the

Fig. 1. Schematic process of the synthesis and preparation of the samples.

36266
K. Albaidani et al. Ceramics International 49 (2023) 36265–36275

Table 1
Results of TiO2 and Cu-TiO2 with various amounts.
Sample No. hkl FWHM (◦ ) Peak pos. [θ◦ ] Crystallite size [Å] d-spacing [Å] Dislocation density[δ] Lattice strain [%]
λ × 10− 5
d =
2 sin θ

1 101 0.5575 12.70 147.8 3.5038 4.4950 0.618


2 101 0.8392 12.75 98.2 3.4903 10.181 0.946
3 101 0.7956 12.80 103.6 3.4769 9.1470 0.875
4 101 0.7524 12.65 109.5 3.5174 8.1880 0.838

space group 141/amd (141) [26]. The XRD diffractograms of films are 3.2. Raman Spectroscopy study
shown in Fig. 2. According to this graph, the positions of the diffraction
peaks of pure TiO2 and Cu-doped TiO2 were the same, and both dis­ This is an important characterization technique for analyzing the
played almost identical shapes. In fact, these peaks correspond to the local structural changes caused by dopant ions. Fig. 3 displays the
TiO2 anatase phases of (101), (004), (200), (105), (211), (204), (116), Raman spectra of pristine and doped TiO2 films, revealing that Raman
and (215) and were found at angles around 2θ = 25.3◦ , 37.8◦ , 48.0◦ , lines are attributed to anatase phase. However, there are no secondary
53.9◦ , 55.1◦ , 62.7◦ , 68.8◦ , and 75.0◦ , respectively. Moreover, the XRD peaks connected to Cu or to any of its oxidized phases. The Eg mode of
diffractograms of Fig. 2 does not show any extra peaks which reveals anatase TiO2 is shown by the sharp peak at 76 cm− 1. One further low-
that no additional species, like copper oxide or other form, were intensity mode (Eg2) is also present, at 120 cm− 1. The Eg3 mode is
discovered. These results are inagreement with. This may be explained presented at about 580 cm− 1. At 330 cm− 1, the B1g -peak arises, and at
by the fact that the Cu species were extensively distributed with few 450 cm− 1, the A1g + B1g -peak appears as shown in Table 2. The Ti-O
TiO2 particles or the number of copper oxides was below the detection stretching and bending vibrations are connected with a distinctive
threshold of the XRD technique [28,33]. According to Table 1, the lat­ Raman vibration. In fact, such O-Ti-O symmetric stretching and bending
tice parameters a and c does not seem to vary much with the Cu content. vibrations are responsible for Eg and B1g modes respectively. While, the
Looking at the ionic radii of Cu2+(0.73 Å) and Ti4+ (0.64 Å), a potential asymmetric bending vibration is responsible for A1g mode. Cu2+ (0.73
replacement of a limited amount of Ti4+ by Cu2+ would be joined by an Å) has a greater ionic size than Ti4+ (0.64 Å), therefore doping this ion
unusually small gate extension. In the higher convergence of the Cu would deform the lattice pattern, as seen in the 0.2Cu-TiO2 result. This is
dopant, CuO can be acquired, but in the XRD structure, there is no due to the charge difference between Cu2+ and Ti4+, and to the oxygen
pinnacle having a place with Cu groups. In addition, it is very significant vacancies created in the lattice to preserve charge neutrality when Cu is
to mention that according to the well-known rules of Hume-Rothery, the added. If doping arises on the substitutional location on a Ti4+ site, the
difference in ionic radius should be less than 15%, which is the case Ti-O-Ti bond will be broken, and a new Cu-O-Ti or Cu-O-Cu bonding will
here; the ionic radius of Cu+2 is almost close to that of Ti4+. This is most occur, causing the peaks to widen and shift. Despite the fact that Cu2+
likely because of the small amounts of the Cu dopant, Cu+2 will be doping on the Ti4+ site affects all modes, we focused on the intense Eg
dissolved in the TiO2 crystal lattice [33,34]. There are low intensity
peaks which may be due to the sodium ions appearing in the film pro­
duced and which are produced at angles about 31.8◦ according to the
reference [2]. The sodium impact could be the result in a phase change
from anatase to brookite [5]. By comparing the results of Raman spec­
troscopy, it can be discovered that there is no transition from the anatase
phase to brookite which indicates that the presence of sodium may be
attributed to the glass substrate [34].

Fig. 3. (a) Raman results of TiO2 and Cu-TiO2 composite samples and (b)
Fig. 2. Diffractograms of TiO2 and a series of Cu-TiO2 thin film samples. Schematic representation of the distortion of the TiO6 octahedra during doping.

36267
K. Albaidani et al. Ceramics International 49 (2023) 36265–36275

Table 2 3.4. Surface morphology characterization


Raman peak position (cm− 1) for the vibration modes Eg1, Eg2, Bg1, Ag1+Bg1 and
Eg3 for pure and doped TiO2. The investigation of morphologies was carried out with SEM, and
Vibration modes Pure TiO2 0.1Cu-TiO2 0.2Cu-TiO2 0.3Cu-TiO2 images of obtained samples can be seen in Fig. 5. All Cu doped TiO2
Eg1 76 76 76 76
samples have agglomerated without any particular morphology. In
Eg2 120 120 120 120 addition, the agglomeration degree is different between the different
Bg1 330 330 330 330 amounts of Cu doped TiO2 [2]. In 0.3Cu-TiO2, the film’s microstructure
Ag1+Bg1 450 450 450 450 (Fig. 5(d)) reveals some agglomeration without any particular
Eg3 580 580 581 582
morphology and cracks throughout the film. Cracking was less extensive
in 0.1Cu-TiO2 and 0.2Cu-TiO2 while 0.3Cu-TiO2 contains very obvious
peak to better know the doping impact. The Cu-O-Cu stretching mode cracks [13], as mentioned earlier.
vibration is connected with the Eg peak, and when doping occurs, the In order to look into the layers’ elemental content, EDS analysis was
intensity of this vibration decreases as oxygen ions form nearby [5]. used and the spectra obtained for pristine and Cu-doped TiO2 are dis­
For Cu-TiO2 composite, the creation of oxygen vacancies near Cu is played in Fig. 5. From this figure, it is clearly noted that the TiO2 sample
theoretically proven. The formation of these oxygen vacancies causes exclusively comprises Ti and O atoms, according to the EDS spectra. Ti,
the lattice to contract and the peak to shift to a larger wavenumber. Cu, and O atoms are present in the doped layers. Two Ti peaks at 0.3 and
Therefore, the displacement and enlargement of the Raman Eg peak can 4.5 keV are also visible in the Cu-TiO2. EDS analysis insured the presence
also be brought on by structural flaws and phonon confinement [5,35]. of Cu in the Cu-TiO2 thin film samples. The maxima of Cu-atoms are
Fig. 3 (b) depicts the breakdown of the Cu-TiO2 lattice and the creation detected at 1.2, 8.0 and 8.9 keV corresponding to the various concen­
of oxygen vacancies (b). trations of Cu which are 0.1, 0.2 and 0.3% respectively. These results are
in accordance with the literature [35]. In the case of silicon, the impurity
3.3. FTIR results might come from the substrate, soil in the air, or the device’s detector,
which could be polluted or old, according to the hypothesized grounds
From Fig. 4, we note a broad absorption band in the range 500-1000 for its presence [42]. According to 2D and 3D AFM images of samples
cm− 1 which is attributed to the vibrational absorption of the Ti–O–Ti that are shown in Fig. 6, the surface morphologies of the doped samples
bonds. While, the Ti-O bond of the anatase phase is responsible for the clearly exhibit Cu as doping atoms. The surface consists of nanoparticles
absorption bands seen at 618 cm− 1 and 570 cm− 1. Furthermore, for the with uniform distribution. While, the values of RMS (square mean plane
absorption at 1063 cm− 1 and the peaks appearing at 1468 cm− 1 are roughness) and Ra (average roughness) were obtained in Table 3. The
assigned to the Ti–O–N bond. The band identified at 2747 cm− 1 might be AFM results suggested that a little amount of Cu doped in the TiO2 could
related to the titanium carboxylate’s asymmetric stretching oscillations. not remarkably change its roughness, while a large quantity of Cu doped
The maxima found within the range 2924–2843 cm− 1 are attributed to in the TiO2 could significantly increase its roughness and change its
the C–H osscillations of the alkane molecules that originate from tita­ morphology. This change could be explained by the doping-induced
nium tetraisopropoxide and 2-propanol, which we used during fabri­ development of surface crystallites with the clustering of smaller
cation. The maximum that formed at 3500 cm− 1 and 1632 cm− 1 grains. The AFM image shows the absence of peculiar surface grain
typically represents a hydroxyl bond and water adsorption, respectively. composition, especially in 0.2 Cu-TiO2 sample (Fig. 6 (c)) the surface is
These peaks demonstrate that the lattice contains hydroxyl radicals. irregular in contrast to Fig. 6 (b) and Fig. 6 (d), which confirms the XRD
Furthermore, no additional peak developed as the copper doping levels result of in terms of crystallization. In addition, sharp formations of
were raised. On the other hand, there is no existing alkoxy group, and various widths and heights were also seen on the surface, as well as grain
the FT-IR results obtained are in good accord with the XRD study boundaries (black regions) and higher clusters (white regions) [13,41,
[35–41]. 42]. These outcomes agree with those from the SEM analysis.

3.5. Surface chemistry characterization

Fig. 7 displays the (Ti-2p) XPS spectra of our synthesized films to


understand shifts and broadening of its binding energy. In fact, no shifts
were noticed and this is similar to that made in the Raman section,
which is consistent with engendering oxygen vacancies. So, the lack of
oxygen atoms causes the Ti4+ state to change to the Ti3+ state. Decon­
volution analysis of the Ti-2p XPS peaks was conducted in order to
investigate the potential creation of Ti3+ sites (see Fig. 7 (c)). Shirley
type background and Gaussian band shape were adopted during the
deconvolution [43]. Fig. 7 reveals that Ti atom in TiO2 only occurs in its
(+4) oxidation state. In addition, the oxidation states (+3) of Ti atoms
are notably higher in 0.1Cu- and 0.2Cu-TiO2 structures and decrease for
0.3Cu-TiO2. This decrease may be attributed to the overlap between the
maxima corresponding to Ti3+ and Ti4+ states to create wider peaks.
Thus, achieving higher amounts of Ti3+ states which overlap with Ti2
p3/2 of Ti4+, could serve to handle widest peaks. It well known that the
XPS data correlates well with the Raman analysis [44], so, XPS spectra of
O-1s are plotted and displayed in Fig. 7 (b). The O1s core level spectrum
fall within the range 530 - 532 eV and exhibits three peaks witch very
close to that found by M. Pérez-González et al. [45]. These distin­
guishable peaks are assigned to lattice oxygen (OL), surface hydroxyl
oxygen (O–OH) and adsorbed oxygen (OS) in TiO2, respectively. Ac­
Fig. 4. FTIR spectra of pure TiO2 and Cu-TiO2 thin film with different amounts. cording to Ref. [46] the XPS could only discover the outermost

36268
K. Albaidani et al. Ceramics International 49 (2023) 36265–36275

Fig. 5. SEM and EDS chemical analysis for (a) TiO2, (b) 0.1Cu-TiO2, (c) 0.2Cu-TiO2, and (d) 0.3Cu-TiO2 thin films.

superficie (≈10 nm) of a sample that explanation somewhat the absence 3.6. TEM study
of (Cu) [47]. This indicates that copper ions may be dispersed at the
bottom of the prepared sample and not at the surface, which is the area Fig. 8(a–d) illustrates results from a TEM test on the detailed mor­
of XPS detector or amount of Cu much smaller below the sensitivity or phologies of TiO2 and Cu-TiO2. Images showed that several shapes may
resolution limit of the XPS which cannot detect it. Tin (Sn) defect was be seen. The image of TiO2 without agglomeration was displayed in
discovered in our results, perhaps the Sn -peaks expand and covered the Fig. 8(a). Fig. 8(b–d) illustrates the heterogeneous distribution of the
copper peaks. coper particles on the surface of the TiO2 [48]. The TEM images clearly
show that the Cu dopant with rectangular shaped structures was well

36269
K. Albaidani et al. Ceramics International 49 (2023) 36265–36275

Fig. 6. 2D and 3D AFM graphs for (a) TiO2, (b) 0.1Cu-TiO2, (c) 0.2Cu-TiO2, and (d) 0.3Cu-TiO2 thin films.

distributed on the TiO2 surface. With the addition of copper to the TiO2
Table 3
matrix, the average particle size reduces, and these findings are
Surface roughness values of the prepared samples.
consistent with the X-ray data. Therefore the presence of copper could
Roughness parameter & Grain TiO2 0.1Cu- 0.2Cu- 0.3Cu- decrease the size of TiO2 crystallites, attributed to the deterioration of
size (nm) TiO2 TiO2 TiO2
the anatase phase of TiO2 [49]. High-resolution transmission electron
Rms 15.43 8.01 14.71 3.47 microscopy (HR-TEM) analysis was also used to study the imaging of
Ra 12.77 6.04 11.41 2.49 TiO2 and Cu doped TiO2 samples. The estimated interplanar spacing of
Grain size 17.86 9.39 13.49 5.00
adjacent lattice fringes of all samples are about 0.349 nm, approaching
closely the 0.352 nm lattice spacing of the anatase (101) planes. These

36270
K. Albaidani et al. Ceramics International 49 (2023) 36265–36275

Fig. 7. (a) XPS survey spectra of the TiO2 and Cu-TiO2 thin films, and high resolution XPS spectra of (b) the O-1s and (c) the Ti-2p3/2 peaks.

Fig. 8. TEM and HRTEM images of the synthesized (a) TiO2, (b) 0.1Cu-TiO2, (c) 0.2Cu-TiO2, and (d) 0.3Cu-TiO2 thin films samples.

lattice distances were in accordance with the X-ray diffraction (XRD) 66.43%, respectively in the spectral range 450-700 nm. The film’s
spectra (Fig. 2). transmittance is influenced by several factors, including poly crystal­
linity, thickness, porosity, and surface roughness [41]. UV–vis spectra of
3.7. Optical properties analysis the prepared samples are displayed in Fig. 9 (b). The absorption of pure
titania is associated to the excitation of the O2p electron to the Ti3d state.
Fig. 9 (a) shows the transmittance (T) spectra for our prepared Cu-TiO2 has an absorption edge extending to larger wavelengths. This
samples. The TiO2, 0.1Cu-TiO2, 0.2Cu-TiO2, and 0.3Cu-TiO2 thin films behavior reveals the substantial interaction between TiO2 and Cu atoms
have average transmittance values of 53.57%, 56.31%, 50.59%, and [49,50]. All spectra shown here exhibit mainly wide absorption band

36271
K. Albaidani et al. Ceramics International 49 (2023) 36265–36275

Fig. 9. (a) Transmittance, (b) absorbance spectra, (c) Tauc plot of TiO2 ln(α) versus (hv) of TiO2 and Cu-TiO2 samples.

within the wavelength range 300–700 nm. This behavior might be disorder is given by the following relation [52];
induced by the dispersion of trap sites and the variation in defect ratios
brought on by oxygen vacancies (as confirmed by Raman and XPS in­ α(hv) = α0 e(hv/Eu ) (4)
vestigations) [51]. Using the following equation, the absorption coeffi­
with α0 being the characteristic parameter of the substance. The Eu of the
cient (α) and the optical band gap energy are determined according to
thin films were tabulated in Table 4. These Eu values are determined by
the Tauc method [41,44,58], as follow;
the slope of the linear portion of the curve in Fig. 9 (d). The Urbach
( )
A 1 1 − R2 energy are 1.94, 1.21, 1.23, and 1.42 eV, respectively for TiO2 and Cu-
α = = ln (2)
d d T TiO2 samples. Hence, as the bandgap reduces, the magnitude of defect
energy increases [41,50]. The extinction coefficient k of the complex
( )2
αE = Aop E − Eg (3) refractive index can be calculated using the relation k = 4απλ . The spectral
variation of the refractive index n is studied using the following equation
where d and Aop are the thickness and optical constant of the layer. [42,53]:
The Eg values were determined from Tauc plot (Fig. 9 (c)) and ( ) √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
summarized in Table 4. From this table it can be noted that Eg increases n=
1+R
+ (
4R
) − k2 (5)
from 3.6 to 3.8eV as the amounts of Cu dopants increases. This may be 1− R 1 − R2
explained by the fact that it might be caused by any charged defects or
The variation of k and n of the samples have been referenced in Fig. 9
that the charged defect generated was neutralized by some others. As a
(a, b) and they are tabulated in Table 4. Fig. 10 (a), shows that the peak
result of the change in bandgap value caused by Cu-doping suggesting
amplitude of the extinction coefficient shrink rapidly within the range
an increase in n-type carrier concentration, most of the Cu ions must be
350-500 nm, especially for 0.2 g and 0.3 g Cu amounts. The inverse of
absorbed into the structure as interstitial donors rather than acceptor
the absorption coefficient δ = 2α = ωck may be used to calculate the skin
replacement. Our findings are consistent with prior research that has
reported adjusting the bandgap energy level of metal oxide thin films via depth (δ) of the samples. The skin depth (δ) is a measurement of the
metal ions doping [42]. optical beam intensity penetration into the material before scattering.
The Urbach’s energy (Eu) which is a result of the material’s structural Because the value is on the order of the nanometer, the accidental laser
beam on the substance barely penetrates a short distance. The skin depth
changes as depicted in Fig. 10 (c). A relational dispersion equation,
Table 4
which specifies the single-oscillator model, is used to describe variations
The optical parameters of the TiO2 and Cu-TiO2 for various Cu amounts.
of material’s refractive index caused by light energy. As a result, the
Sample Eg Eu Ed E0 n0 ε͚ k n refractive index dispersion relation is provided following to Ref. [41], by
(eV) (eV) (eV) (eV) (eV)
the following relation;
TiO2 3.60 1.94 1.602 4.58 1.16 2.60 0.043 0.276
0.1Cu- 3.73 1.21 1.29 4.60 1.14 3.25 0.029 0.161
( )− 1 E0 1
n2 − 1 = − (hv)2 (6)
TiO2 Ed E0 Ed
0.2Cu- 3.74 1.23 1.519 4.57 1.13 3.08 0.034 0.298
TiO2 Here Ed is the dispersion energy, and E0 is the energy of a single
0.3Cu- 3.81 1.42 1.35 4.30 1.13 3.53 0.02 0.161 oscillator. The dispersion relation of Eq. (6) is computed and displayed
TiO2 in Fig. 10 (d). The values of E0 and Ed were calculated from the intercept

36272
K. Albaidani et al. Ceramics International 49 (2023) 36265–36275

Fig. 10. (a) Change of extinction coefficient, k, (b) refractive index, n, (c) Change of δ versus hv and (d) variation of (n2-1)− 1
versus (hv)2 for TiO2 and doped
TiO2 samples.

Fig. 11. Variation of the real (a), the imaginary (b) parts of the dielectric permittivity (εr and εi) with the wavelength, and variation (c) of the real part of dielectric
permittivity (εr) with λ2.

36273
K. Albaidani et al. Ceramics International 49 (2023) 36265–36275

of the extrapolation to the (n2 − 1)− 1 axis and the slope, respectively. [55] was considered. In fact, the DFT + U approach combine the on-site
Table 4 lists the dispersion and the single oscillator energy of our thin intra-atomic electron-electron interaction (U) and the exchange corre­
film samples. Similarly, the static refractive index n0 was calculated by lation (J) contributions into a unique effective U-parameter to estimate
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ the exchange− correlation energies between Ti-3d and O-2p orbitals
n0 = 1 + Ed/E . and results were given in Table 4. The values of (εr)
0 leading to a more optimal band gap. To determine the suitable
and (εi) have been studied from (n) and (k) values as in Refs. [41,54]: U-parameter to precisely understand the intra-atomic electron correla­
εr = n2 − k2 (7) tion, we plot the relationship between U and Eg for pure anatase TiO2
and the results are shown in Fig. 13 (a). This figure reveals that the band
εi = 2nk (8) gap broadens by increasing the effective Hubbard parameter of Ti. Based
on the same figure, one could notice that the optimized band gap value
In Fig. 11 (a, b) these values have been plotted as a function of the
wavelength, the real part of the dielectric coefficient can be computed
using the following relation [41]:
( )
Ne2
εr = ε∞ − λ2 (9)
4π 2 C2 ε0 m∗

here ε∞ is the material’s permittivity at high-frequencies, e is the elec­


tron charge, c is the vacuum light-speed, ε0 is the free space permittivity,
Ne is the free carrier density and m* is the effective mass. The value of ε∞
can be determined by drawing a straight line at higher wavelengths on
the curve εr = f(λ2). If εr versus λ2 graph is made, a straight line may be
drawn at higher wavelengths, allowing the high-frequency dielectric
constant to be calculated. In Fig. 11 (c), the changes of εr with λ2 have
been displayed and the calculated values of ε∞ are 2.60 eV, 3.22 eV, 3.08
and 3.53 eV for TiO2, 0.1Cu-TiO2, 0.2Cu-TiO2, and 0.3Cu-TiO2 thin films
respectively.

4. Theoretical analysis

To understand the increase of the gap energy obtained experimen­


tally for our Cu-doped TiO2 structures and analyze the effects of Cu
content on the change in the band gap, we performed theoretical cal­
culations. All our theoretical investigations are within the framework of
quantum-mechanical procedure, employing density functional theory
(DFT) computations. Here, we adopted DFT with Hubbard-based U
adjustment, specifically, the DFT + U approach to extract the energy gap
of the pure as well as doped TiO2 films. Our DFT computations have
been accomplished employing quantum espresso ab initio software [55,
56]. To model the exchange-correlation energy in our calculations,
generalized gradient approximation (GGA-PBE) scheme has been
adopted [57].
First, we have calculated the electronic band diagram and DOS of
pristine anatase TiO2 and the results are reported in Fig. 12. According
to this figure, one can easily notice that the energy band gap is about
2.16 eV. This calculated value is significantly lower than the experi­
mental one which is about 3.6 eV. This underestimated band gap value Fig. 13. (a) Relationship between the Hubbard U and Eg for anatase TiO2, and
could be assigned to the selection of exchange-correlation energy. So, to (b) DOS plots calculated for undoped anatase TiO2 and Cu-TiO2 for
recover the precision in band gap computations, the Hubbard U model different amounts.

Fig. 12. Calculated electronic band structure and DOS of anatase TiO2.

36274
K. Albaidani et al. Ceramics International 49 (2023) 36265–36275

which agree the experimental value (Eg = 3.58 eV), is obtained for U = 8 [11] N.D.S. Mohallem, M.M. Viana, M.M.L. de Jesus, G.H. de Magalhães Gomes, L.F. de
Sousa Lima, and E.D.L. Alves. London (2018)..
eV. To examine the impact of the fraction of Cu dopant on the energy gap
[12] A. Timoumi, Graphene 7 (4) (2018) 31–38.
of the Cu-doped TiO2 structures, DOS are calculated for various Cu [13] R. Vidhya, R. Gandhimathi, M. Sankareswari, P. Malliga, J. Jeya, K. Neivasagam,
concentrations and the results are shown in Fig. 13 (b). It reveals that as J. Nanostruct. 8 (3) (2018) 232–241.
the extend of Cu doping get larger, the valence band maximum is low­ [14] D. Jiang, Y. Ouyang, T. Otitoju, Y. Ouyang, N. Shoparwe, D. Zhang, S. Wang,
A. Zhangand, S. Li, Catalysts 11 (2021) 1039.
ered, and the conduction band edge is raised leading to an excess in band [15] J. Wang, X. Li, Y. Ren, Z. Xia, H. Wang, W. Jiang, C. Liu, S. Zhang, Z. Li, S. Wu,
gap energy. Our calculated Eg values are 3.58 eV, 3.7 eV, 3.77 eV and N. Wang, G. Liu, S. Liu, W. Ding, Z. Zhang, J. Alloys Compd. 25 (2021), 157726.
3.91 eV corresponding to the structures; pure-TiO2, 0.1Cu-TiO2, [16] G. Nagaraj, M. Mohammed, H. Abdulzahra, P. Sasikumar, S. Karthikeyan,
S. Tamilarasu, Appl. Phys. A (2021) 127–269.
0.2Cu-TiO2 and 0.3Cu-TiO2 respectively. These energy values are [17] T. Rajaramanan, S. Shanmugaratnam, V. Gurunanthanan, S. Yohi,
consistent with our experimental results. D. Velauthapillai, P. Ravirajan, M. Senthilnanthanan, Catalysts 11 (2021) 690.
[18] M. Abbas, M. Rasheed, J. Phys. Conf. 1795 (2021), 012059.
[19] F.S.S. Zahid, M.S.P. Sarah, M.Z. Musa, U.M. Noor, M. Rusop, IEEE International
5. Conclusions Conference on Electronics Design, Systems and Applications (ICEDSA) (2012)
22–26.
In conclusion, TiO2 and Cu-doped TiO2 thin films have been suc­ [20] A. Kösemen, Z.A. Kösemen, B. Canimkubey, M. Erkovan, F. Başarir, S. ErenSan,
Osman Örnek, Ali Veysel Tunç, Sol. Energy 132 (2016) 511–517.
cessfully fabricated and analyzed using several techniques. The XRD [21] R.S. Ganesh, AamirY. Mamajiwala, E. Durgadevi, M. Navaneethan, S. Ponnusamy,
spectra have shown the formation of the anatase TiO2 phase with pref­ C.Y. Kong, C. Muthamizhchelvan, Y. Shimura, Y. Hayakawa, Mater. Chem. Phys.
erential orientation (010). The crystallite size was expanded from 108 to 234 (2019) 259–267.
[22] Y. Lva, L. Yua, H. Huang, H. Liu, Y. Feng, J. Alloys Compd. 488 (2009) 314–319.
113 Å upon incorporating Cu. Shift with doping in the TiO2 lattice as [23] A. Timoumi, H.M. Albetran, H.R. Alamri, S.N. Alamri, I.M. Low, Superlattice.
well as a specific absorption band between 400 and 1000 cm− 1 have Microst. 139 (2020), 106423.
been demonstrated based on Raman spectroscopy analysis. In addition, [24] W. Belhadj, A. Timoumi, F.A. Alamer, O.H. Alsalmi, S.N. Alamri, Results Phys. 30
(2021), 104876.
spherical shape of the nanoparticles with a high degree of agglomeration [25] W. Belhadj, A. Timoumi, H. Dakhlaoui, F.A. Alamer, Coatings 12 (2022) 129.
and homogeneity was visualized by the SEM technique, with d-spacings [26] Y. Liu, H. Zhou, J. Li, H. Chen, D. Li, B. Zhou, W. Cai, Nano-Micro Lett. 2 (2010)
of 3.05 and 1.88 Å recorded for the pure and spiked sample, respec­ 277–284.
[27] A. Timoumi, S.N. Alamri, H. Alamri, Results Phys. 11 (2018) 46–51.
tively. The TEM images show that the particle size decreases with
[28] A. Timoumi, W. Zayoud, A. Sharma, M. Kraini, N. Bouguila, A. Hakamy,
addition of copper which was distributed on the TiO2 surface. Moreover, N. Revaprasadu, S. Alaya, J. Mater. Sci. Mater. Electron. 16 (2020) 13636–13645.
our optical investigations showed an absorption tail that trailed into the [29] S. Rtimi, C. Pulgarin, J. Kiwi, Coatings 7 (20) (2017).
wavelength range with increasing dopant amounts, enabling the [30] A. Timoumi, H. Bouzouita, Int. J. Renew. Energy Technol. Res. 2 (7) (2013)
188–195.
bandgap to increase from 3.6 to 3.8 eV. Finally, theoretical study using [31] A. Timoumi, M.A. Wederni, N. Bouguila, B. Jamoussi, M.K. Alturkestani,
the DFT + U approach reveals that the conduction band edge is raised R. Chakroun, B. Al-Mur, Synth. Met. 272 (2021), 116659.
leading to an increase in the band gap energy. The material is suited for [32] A. Timoumi, M.K. Al Turkestani, S.N. Alamri, H. Alamri, J. Ouerfelli, B. Jamoussi,
J. Mater. Sci. Mater. Electron. 27 (2017) 7480–7488.
upcoming photovoltaic applications incorporating various architec­ [33] F. Bensouici, M. Bououdina, A.A. Dakhel, R. Tala-Ighil, M. Tounane, A. Iratni,
tures, such as thin film solar cell devices, thanks to the overall features. T. Souier, S. Liu, W. Cai, Appl. Surf. Sci. 395 (2017) 110–116.
[34] A. Tamarani, R. Zainul, I. Dewata. J. of Physics: Conf. Series 1185 (2019), 012-020.
[35] B.R. Sankapal, M. Ch Lux-Steiner, A. Ennaoui, Appl. Surf. Sci. 239 (2005) 165–170.
Declaration of competing interest [36] B. Choudhury, M. Dey, A. Choudhury, Int. Nano Lett. 3 (25) (2013).
[37] S. Al-Taweel, R. Saud, J. Chem. Pharmaceut. Res. 8 (2016) 620–626.
[38] L. Chougala, M. Yatnatti, R. Linganagoudar, R. Kamble, J. Kadadevarmath,
The authors declare that they have no known competing financial J. Nano- Electron. Phys. 9 (2017), 404-005.
interests or personal relationships that could have appeared to influence [39] B. Rajamannan, S. Mugundan, G. Viruthagiri, P. Praveen, N. Shanmugam,
Spectrochim. Acta Mol. Biomol. Spectrosc. (2014) 651–656.
the work reported in this paper.
[40] M. Yıldırım, J. Alloys Compd. 773 (2019) 890–904.
[41] N.A. Kattan, S.A. Rouf, N. Sfina, M. mana Al-Anazy, H. Ullah, A. Hakamy, A. Mera,
Acknowledgments Q. Mahmood, M.A. Amin, J. Solid State Chem. 319 (2023), 123820.
[42] M. Ahamed, M.A. Majeed Khan, M.J. Akhtar, H.A. Alhadlaq, A. Alshamsan, Sci.
Rep. 6 (2016), 30196.
The authors extend their appreciation to the Deputyship for Research [43] M. Morales-Luna, S.A. Tomás, M.A. Arvizu, M. Pérez-González, E. Campos-
& Innovation, Ministry of Education in Saudi Arabia for funding this Gonzalez, J. Alloys Compd. 722 (2017) 938–945.
[44] A.M. Abd-Elnaiem, R.M. Hassan, H.R. Alamri, H.S. Assaedi, J. Mater. Sci. Mater.
research work through the project number: IFP22UQU4331172DSR150.
Electron. 31 (2020) 13204–13218.
[45] M. Pérez-Gonzáleza, S.A. Tomás, Catal. Today 360 (2021) 129–137.
References [46] R. Ako, P. Ekanayakea, D. Younga, J. Hobley, V. Chellappan, A. Tan, Gorelik,
G. Subramanian, Ch Lim, Appl. Surf. Sci. 351 (2015) 950–961.
[47] B. Xin, P. Wang, D. Ding, J. Liu, Z. Ren, H. Fu, Appl. Surf. Sci. 254 (2008)
[1] J. Tao, X. Hu, J. Xue, Y. Wang, G. Weng, S. Chen, Z. Zhu, J. Chu, Sol. Energy Mater.
2569–2574.
Sol. Cell. 197 (2019) 1–6.
[48] B. Moongraksathum, J.-Y. Shang, Y.-W. Chen, Catalysts 8 (2018) 352.
[2] L. Al-Salem, R. Seoudi, J. Mater. Sci. Mater. Electron. 31 (2020) 22642–22651.
[49] S. Rtimi, C. Pulgarin, J. Kiwi, Coatings 7 (2017) 20.
[3] B. Ünlüa, M. Özacar, Sol. Energy 196 (2020) 448–456.
[50] I.H. Tseng, J.C.S. Wu, H. Ying Chou, J. Catal. 221 (2004) 432–440.
[4] L. Ryan, Department of Physics, Oregon State University, 2018.
[51] B. Choudhury, B. Borah, A. Choudhury, Photochem. Photobiol. 88 (2012)
[5] A. Alotaibi, B. Williamson, S. Sathasivam, A. Kafizas, M. Alqahtani, C. Sotelo-
257–264.
Vazquez, J. Buckeridge, J. Wu, S. Nair, D. Scanlon, I. Parkin, ACS Appl. Mater.
[52] F. Urbach, Phys. Rev. 92 (1953) 1324.
Interfaces 12 (13) (2020) 15348–15361.
[53] J.R. Mohamed, L. Amalraj, Journal of Asian Ceramic Societies 4 (3) (2016)
[6] M. Liu, L. Zheng, X. Bao, Z. Wang, P. Wang, Y. Liu, H. Cheng, Y. Dai, B. Huang,
357–366.
Z. Zheng, Chem. Eng. J. 405 (2021), 126654.
[54] A.M. Abd-Elnaiem, et al., J. Mater. Sci. Mater. Electron. 33 (2022) 23293–23305.
[7] Y. Zhang, W. Chu, Chemosphere 286 (2) (2022), 131797.
[55] P. Giannozzi, et al., J. Phys. Condens. Matter 21 (2009), 395502.
[8] P. Narayanama, S. Major, Colloids Surf. A Physicochem. Eng. Asp. 609 (2021),
[56] P. Giannozzi, et al., J. Phys. Condens. Matter 29 (2017), 465901.
125652.
[57] V.I. Anisimov, J. Zaanen, O.K. Andersen, Phys. Rev. B Condens. Matter 44 (1991)
[9] R. Katal, S.M. Panah, M. Tanhaei, M. Hossein, D.A. Farahani, H. Jiangyonga, Chem.
943–954.
Eng. J. 384 (2020) 1385–8947.
[58] A.S. Hassanien, H.R. Alamri, I.M. El Radaf, Opt. Quant. Electron. 52 (2020) 335.
[10] L. Pedrinia, L. Escaliantea, L. Scalvia, Mater. Res. 24 (2021).

36275

You might also like