You are on page 1of 13

Applied Surface Science 536 (2021) 147843

Contents lists available at ScienceDirect

Applied Surface Science


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

Highlighting the importance of optimal defect density through band T


structure and photocatalytic studies
Dileep Maarisetty , Janaki Komanduru, Saroj Sundar Baral

Department of Chemical Engineering, BITS Pilani, K K Birla Goa Campus, Goa 403726, India

ARTICLE INFO ABSTRACT

Keywords: Interface energetics plays a substantial role in the efficient charge transfer between semiconductor materials.
Dipoles The concentration of donors or acceptors at the junction determines the direction of charge flow. In this work,
Fermi level CdS was doped to TiO2 progressively to explore the electronic structure modification and its associated effect on
Work function charge dynamics. Results suggest that loss of dipole moment with increasing CdS concentration; causing the
Photocatalytic degradation
work function (WF) values of the composites to increase. The dipole-dipole repulsion occurs as a consequence of
Oxygen reduction reaction
Optoelectronics
a decrease in Cd-Cd distance at higher doping concentrations. When present in optimal concentration, the
surface dipoles besides aiding in efficient charge transfer, accelerate the chemisorption process at corrugated
surfaces. Apart from the optoelectronics, efforts have been made to understand the hindering role of higher
defect concentration in triggering the thermodynamically favored undesired photocatalytic reactions, which
drastically reduce the quantum yield of the process. This work may provide some insights to tune the work
function (WF) values in applications that involve optoelectronic and photocatalytic studies.

1. Introduction modulating the interfacial properties (favoring the dissociative ad-


sorption), electronic properties and regulating atom coordination
Reports from photocatalysis research prove the inconsistency in number as understood from theoretical studies [5]. Recently an ex-
terms of quantum yield, even when the same composition of photo- tension in photocatalytic applications such as CO2 reduction [6], or-
catalysts are compared. A huge number of variables influencing it such ganic synthesis [7] and nitrogen fixation [8] have taken a new leap in
as type of precursors, synthesis method, doping concentration and also defect engineering and achieved tremendous results. So far many re-
heating treatments lead to varying results. The major contributor con- searchers have worked on CdS-TiO2 as photocatalysts and confirmed
undrum continues but emphasizes the importance of looking this pro- their superior performance than their individual photocatalytic per-
blem at a molecular level. Takanabe et al., reported that defect density, formance under visible light [9]. Numerous studies have reported the
charge carrier concentration and interfacial properties as three strong varying concentration of dopants and its effect on bandgap [10], re-
influencing candidates for enhancing the photocatalytic efficiencies [1]. combination and photocatalytic activity [11] yet the underlying rea-
Interestingly, carrier concentration and interfacial properties are con- sons were not fully studied [12]. A systematic approach to this can be
trolled by the defects thereby highlighting the key role of surface de- achieved by considering the acceptor or donor levels at the interface/
fects in photocatalytsis. Our previous work served as a complementary junction (attributed to doping) between two semiconductors that dic-
to this, where defect-rich catalyst exhibited higher photocatalytic ac- tate the electron transfer [13]. The decrease in work function values,
tivity compared to the same composite (but with lesser defects), in spite particularly, at the interface between two metal oxides, can be reasoned
of having 6–7 times higher surface area than the former [2]. Recently, a by band bending due to the surface dipole moments; induced by de-
review article elaborately explained the role of defects in boosting the fects. However, this is not straight forward as one may think, and
photocatalytic activity on a fundamental level and readers are en- deeper understanding of the relationship between dopant concentra-
couraged to go through it [3]. Although strategies like facet, phase, tions and work function is required to come to a proper conclusion
morphology control, charge transfer schemes were reported [4], defect [14].
engineering serves as an effective plan of action for enhancing the Also, chemical and structural variations can account for the change
photocatalytic efficiency. Defect-engineering displays potential in in work function values such as transformation from Ti+4 to Ti+3 [15].


Corresponding author.
E-mail address: dileepnaidu59@gmail.com (D. Maarisetty).

https://doi.org/10.1016/j.apsusc.2020.147843
Received 2 March 2020; Received in revised form 24 August 2020; Accepted 5 September 2020
Available online 11 September 2020
0169-4332/ © 2020 Elsevier B.V. All rights reserved.
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

a b

c d
CdS Anatase (101)

TiO2

Fig. 1. (a) XRD patterns of composite catalysts (b) Crystallite size changes in composites after doping CdS (c) TEM image depicting Cd clusters on TiO2 nano particles
(R4) (d) SAED image of R4 confirming crystallinity (no lattice distortion) after CdS doping.

Few reports also claim the underperformance of photocatalysts due to 2. Materials and methods
excessive defect density by creating more recombination centers in the
forbidden gap [3]. An optimized concentration of these surface defects 2.1. Material synthesis
is very important, which would otherwise favor the undesired back-
ward reactions and account for decreased photocatalytic efficiency All chemicals used during the preparation were of analytical grade.
[16]. Recently, multi-component ZnxCd1−xS and ZnIn2S4 have received Ti(OBu)4 and CdSO4 were used as precursors for TiO2-CdS composites
considerable attention recently in the field of photocatalysis owing to where ethanol served as a solvent. Appropriate amounts of Ti(OBu)4
moderate bandgaps [17]. An efficient separation of photo-excited and CdSO4 were added dropwise and stirred for 1 h. The reaction
charge carriers by surface dipoles was reported due to the sulfur va- system was transferred to a hydrothermal reactor, which was operated
cancies and interstitials [18], the readers are recommended to go at 200 °C. The precipitate was rinsed with water and dried in the oven
through it.. Although co-catalysts decrease the energy offset to avoid at 80 °C. The obtained product was heated at 450 0C for 3 h to remove
recombination, higher dopant concentrations enhance the defect den- the organics and improve the crystallization of the material. The com-
sity. A controlled amount of S-vacancies for optimum photocatalytic posites were synthesized at different Cd/Ti weight ratios by following
performance was reported where ZnS structures exhibited a better identical procedure. Hereafter, the composites with varying Cd/Ti ra-
photo-stability [19]. In this work, dipoles resulted from the defects tios (weight) such as 0.1, 0.3, 0.5 and 0.7 are denoted as R1, R2, R3 and
were comprehensively studied at the heterojunction; as dipole moment R4 respectively.
ensures smooth flow of electrons. With Photocatalysis having numerous
parameters that decide the overall efficiency, Takanabe in his work 2.2. Photocatalytic experiments
emphasized the importance of interface properties along with carrier
concentration, mass transfer and defect density [1]. Interestingly, three The photocatalytic activity was tested in a 100 ml reaction system
out of four aspects (excepting mass transfer) in the above mentioned (20 ppm MB + catalyst). To ensure adsorption–desorption equilibrium,
aspects form connection to defects. Hence, it can be understood that 0.05 g catalyst was dispersed into a 20 ppm MB and allowed to stir for
defect engineering holds an important place in modifying the optoe- 1 h under dark conditions. Followed by this, photocatalytic experiments
lectronic properties, interfacial contact and dissociative study [3]. By were conducted under sunlight. Samples were drawn from the reaction
considering their importance, defect induced surface dipoles and their system at periodic intervals (30 min) for further analysis. The con-
consequential effect on surface potentials investigated by systematically centration of MB was determined using UV–Vis spectrophotometer (UV-
increasing their concentration and validating through various char- 3000, Lab India) at an absorption wavelength of 668 nm.
acterization tools and photocatalytic studies.
2.3. Characterization

Morphological studies were carried out by using the field emission

2
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

scanning electron microscope (FESEM). Bandgap studies were con-


ducted using UV–visible diffuse reflectance spectra (UV-DRS) in a wa-
velength range of 200–800 nm (Shimadzu, Japan). The extent of charge
recombination after photo-excitation (ƛ=325 nm) was investigated
with photoluminescence (PL) spectra, obtained from fluorescence
spectrophotometer (Perkin Elmer-LS-55). Surface properties and work
function of the catalysts were determined by X-ray photoelectron
spectroscopy (XPS) and ultraviolet photoelectron spectroscopy (UPS)
respectively in PHI 5000 Versa Probe-II (K alpha from Thermo Fischer
Scientific instruments, UK). For XPS analysis, samples were initially
dispersed in ethanol and drop casted on a silicon wafer. In the next step,
the sample was loaded in a lock chamber where vacuum was main-
tained until the value reached 10−7 mbar and then shifted to the
analysis chamber. C 1 s peak was taken as a reference point and flood
gun was employed to neutralize the charge effect. Electron spin re-
sonance (ESR) spectra were recorded using ESR-JEOL Japan (X-band).

3. Results and discussion


Fig.2. Recombination study at an excitation wavelength of 532 nm (PL
3.1. Impact of crystallographic modifications on the electronic structure spectra).

The change in optoelectronic properties and charge transfer be-


tween metal oxide and metal sulfide in a composite are dependent on
crystallographic modifications resulting from the dopant atoms. Peaks
at 28.20, 24.80, and 47.9° (Fig. 1a) correspond to (1 0 1), (1 0 0), and
(1 0 3) planes of CdS whereas 25.20, 37.80, 54.10 are associated to the
(1 0 1), (1 0 3), (1 0 5) planes of anatase TiO2 respectively [20]. Fig. 1b
shows the drastic change in crystallite size of composites in comparison
to pristine TiO2. This decrease in intensity and simultaneous peak
broadening suggest that Cd clusters are restricted (Fig. 1c) at grain
boundaries and result in inhibition of crystal growth [2]. Here, it should
be noted that since Cd+2 ionic radius (0.097 nm) is too big to enter into
Ti+4 (0.068 nm) lattice, the possibility of Cd+2 entry is ruled out. This
can be confirmed from SAED (selected area electron diffraction) images
of R4 (highest Cd doped sample) where CdS doping did not account for
any lattice distortions or amorphous nature in TiO2 lattice. However, Ti
(IV) ions penetrated into Cd(II) lattice to account for drastic changes in
the absorption spectra and electronic properties of the. In this case, it
should be noted that Cd being group-II acceptor becomes the host atom.
This ensures a positive charge on Ti and acts as electron traps. These Ti
ions are neutralized by electron entrapment around its vicinity by at-
tractive coulombic potential well.
These vacancies and traps prolongs the lifetime of electrons and
holes [21]. Owing to lower energy requirement for electron excitation
from defect states to conduction band, enhanced light absorption can be
expected in the composites (absorption tail up-shift in visible and NIR).
PL spectra of the samples were recorded for TiO2, CdS and the com-
posites (of different compositions). As observed in Fig. 2, the band-band
recombination in the composites has drastically decreased due to
charge transfer between the catalysts. Since photocatalytic experiments
were conducted under sunlight, the excitation was done in the visible
range (532 nm). R2 exhibits the least recombination due to efficient
charge separation while the increased PL intensity at higher dopant
concentrations is attributed to resistance to charge flow at the junction
that ultimately decrease the lifetime of charge carriers; favoring band-
band recombination [22].
Fig.3. (a) UPS spectra of catalysts, (b) Variation of work function values with
doping concentration.
3.2. Dipole moment and dopant concentration

Furthermore, UPS measurements were conducted to determine the 7.88 eV (Fig. 3(a)) respectively and the variation of WF with CdS
WF values of as-synthesized catalysts. From the experimental results doping can be understood from Fig. 3(b). Generally, whenever two
obtained in this study, the work function (WF) values of TiO2 and CdS dissimilar semiconductors are brought into contact, an electric built-in
are 6.1 and 3.43 eV respectively. The composites behaved differently as field (due to WF difference) is formed at the junction causing the
compared to their pure compounds and opened an avenue for the dis- electron flow from one to another till their Fermi levels are equilibrated
cussion into material aspects. The WF values of R1, R2, R3 and R4 cal- [24]. Since the Fermi level of CdS is higher (WF = 3.43 eV), the
culated from equation ϕs = hυ–Esc [23] are 4.96, 4.86, 6.42 and electron flows from CdS to TiO2 until a new equilibrium value is

3
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

established between them. Typically, the new equilibrium value that et al., reported the reduction in intermolecular distance causes depo-
was attained from this charge transfer falls between 3.43 and 6.1 eV. As larization (loss of dipoles)[28]. This indicates that the dipole moment of
a result, CdS acquires positive charge while TiO2 forms a negative CdS is greatly reduced when it aggregates on the surface of TiO2. In
charge on its side. The differences in charges at the interface cause a another report, the decrease in WF values with an increase in coverage
band bending, which additionally hinders the electron flow from CB of layer was linked to electron affinity enhancement [27]. This was rea-
CdS to TiO2 CB. This coulombic repulsion prolongs the carrier life time soned by increased dipole-dipole repulsion with aggregation causing a
[25]; as reflected from PL spectra (Fig. 2). It is noteworthy to under- decrease of its surface energy by electrostatic depolarization. However,
stand that this mode of charge transfer (TiO2 CB to CdS VB) not only both reports are supporting our results where SEM images clearly depict
provides the spatial separation of charge carriers but also drives the the aggregates of CdS at higher concentrations. Although this effect is
photocatalytic reactions due to the enhanced redox potentials. Few negligible (as long as atoms are strongly bound), the reduction in dipole
reports claim the initial decrease and sudden increase of WF values in moment was observed via other mechanisms due to the interaction
other composites [1326]. However, there are no explanations why between dipoles [29]. Hence, interaction between dipoles at higher
these materials behave in such a way. defect densities affects the electrical conductivity of composites and
The most striking consideration about WF is the discontinuity of the increase the recombination of electron-hole pair.
surface that causes atomic reconstruction (charge redistribution) to give The schematic representation of the change in WF values with di-
rise to two different situations. First, electron spilling normally to the pole concentration is seen in Fig. 6. In Fig. 6, Rm accounts for R1 and R2
surface to form a negative dipole layer. Second, the contraction of while Rn represents R3 and R4 respectively. The work function values
electron charge clouds occurs to produce a positive dipole [15]. As (WF) of TiO2, R1, R2, R3, R4 and CdS are 6.1, 4.96, 4.86, 6.42, 7.88, and
expected, the first case increases the WF while the second case de- 3.43 eV respectively. The modulation of WF values majorly depends on
creases. Here, the initial decrease in the work function value of the the dipoles existing at/around the interface. The dipoles when present
composite is due to the charge transfer from Cd atoms to the substrate in excess concentrations cause a sudden change in the electrostatic
Ti. Followed by this, the Cd atoms acquire a positive charge and repel potential at the interface [28]. This can be observed as up-shift in the
one another to have a uniform distribution on the Ti surface. In short, vacuum level in Rn (R3 and R4) (Fig. 6). The dipoles usually possess
each Cd atom becomes an electric dipole directed outward to bring the direction from negative to positive. When the interaction between di-
WF down. poles increases (in case of Rn), its direction is reversed thereby causing a
resistance to the electron flow. Furtherly, the increase in defect con-
3.3. Fermi level pinning centration gives rise to Fermi level pinning (FLP). The synergistic effect
of modulation of vacuum level (due to dipoles) and FLP caused an
Another interlinking factor to surface dipoles that control the elec- abrupt increase in the WF values of R3 and R4. On the contrary, for R1
trical conductivity is the Fermi level pinning effect (FLP). Tak et al. and R2, the surface dipoles accelerated the electron transfer to the
reported the FLP at higher point defect concentrations to increase the surface and resulted in lowering of work function values. These de-
WF values at extreme conditions i.e., oxygen-deficient and oxygen rich creased WF values were instrumental in promoting the Z-scheme mode
environment (Ga vacancies) [13]. This FLP was witnessed as increased of electron transfer in TiO2-CdS composites.
work function values in R3 and R4, which could be ascribed to two Bandgap studies reveal the role of CdS concentration in enhancing
factors: excess density of point defects and the loss of dipole moment; the absorption spectra of the composites. Usually the bands of different
nonetheless, the latter is the result of former. Fig. 4 (a-f) represent the energies are created by the overlapping of orbitals between neighboring
VBM tails shifts with oxygen vacancies. Since R2 possess the least WF atoms. In TiO2, the conduction in particular arise from the 3d orbitals,
value, it should be noted that smooth flow of electrons is ensured at must transfer electrons to the surrounding atoms. The valence band and
higher optimum defect concentration (Fig. 3b). As the vacancy con- conduction band tends to bend at the heterojunction and behave dif-
centration is increased, the WF values is increased beyond TiO2 value, ferently. The extent of band bending in composite increased with the
and confirming the resistance to the electron flow at higher defect CdS concentration due to higher electron transfer before the loss of
concentration. To further support this statement about FLP, Cd 3d dipole moment. The bandgap values of the composites decreased with
spectra was plotted. As well known before, the increase in binding the CdS concentration (see Fig. 7) and are consistent with other reports
energies corresponds to loss of electrons attributed to the extra elec- [30,31]. Kulbeka-Munk equation was used to estimate the bandgap of
trons (induced by S-vacancies). Fig. 4g depicts the reverse flow of catalysts. The values showed consistent decreasing trend with CdS
electrons in R3 and R4 from TiO2 to CdS, where the binding energy concentration such as TiO2 (3 eV), R1 (2.39 eV). R2 (2.34 eV), R3
shifts were towards the lower energy; indicating the interaction of di- (2.21 eV), and R4 (2.17 eV) (Fig. 7). Similar results attributed to doping
poles at higher vacancy concentration. Fig. 4h, i depicts the schematic were reported recently; where Cd doping decreased the bandgap of
diagram of defect in the forbidden gap and subsequent band edge TiO2-CdS composites [32,33]. Another interesting observation in the
modulations ascribed to FLP. study is the band edge modulation. Here, the S-vacancies accelerates
This band bending was also confirmed in Fig. 4(a–f) in the form of the electron transfer, arrests recombination and thereby boosting the
VBM tail stretching at the interface. Fig. 5(a) and (b) represent the SEM photocatalytic activity [34]. The S-vacancies are deep donors present at
images of CdS and TiO2 respectively. In composites (Fig. 5(c–e)), the 1.72 eV above the VBM of CdS [35]. Here, it should be understood that
CdS atoms are decorated on the TiO2 surface. However, at lower doping both S-vacancies and oxygen vacancies (created due to interaction be-
concentrations, the morphological deformation (Fig. 5(c, d)) is subtle. tween titanate and sulfanate) have donor states below conduction band
With the further increase in doping concentration, CdS atoms agglom- and may cause the up-shift of conduction band minima. As the CdS
erate on the surface and hinder the growth of crystallites [2]. As a concentration is increased, the oxygen vacancies also increase and re-
consequence, the dipoles are more densely packed (Fig. 5(e, f)) to have sult in up-shift of VBM tails [36] as observed in the Fig. 4(a–f). The
more interaction between them and consequently lose their strength modification of both VBM and CBM tails owing to oxygen defects was
[27]. When this interaction becomes significantly higher, it leads to widely discussed in black TiO2 [37,38]. A similar effect was demon-
depolarization effect causing the negative charge to flow back to CdS. strated in Fig. 6 where VB was up-shifted with CdS concentration. In
This increases the WF value of the composites (R3 and R4) higher than another study the raise of CBM and VBM was reasoned to surface dipole
TiO2; as observed from Fig. 3a. Moreover, the depolarization caused the moment, stemming from oxygen vacancies [39]. The dipoles at the
Fermi level to shift in reverse direction and thus leading to FLP at surface orient in such a way that their electrostatic potential is raised
higher defect densities. Also, the reduced intermolecular distance be- (relative to vacuum). This rise in band potentials have particular im-
tween Cd atoms gives rise to high interaction between dipoles. D. Cornil portance in decreasing the schottky barriers. In conclusion, the band

4
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

Fig. 4. (a–f) VB spectra of TiO2, CdS, R1, R2, R3, R4 respectively. (g) Cd 3d spectra illustrating back flow of electrons with sudden shift in binding energies (h)
Schematic diagram of defect formation and subsequent modulation of Fermi level, CB and carrier concentration (i) Defect induced dipoles.

edges are highly sensitive to the vacancies, and their controls are ne- 45 meV [44] respectively; indicating the closeness to Fermi level po-
cessary to obtain desired energy position to drive the photocatalytic sition. When these S-vacancies are present in excess amount, it leads to
degradation. lack of availability of charge carriers for driving the photocatalytic
reactions. On the other hand, the higher defect density triggers the
3.4. Mode of electron transfer and mechanism of photocatalytic reactions thermodynamically favorable undesired back reactions to decrease the
overall yield of the process (will be explained later).
Generally, the XPS spectra of Ti 2p, Cd 3d and S 2p display doublet Ti 2p and Cd 3d spectra confirmed the electron transfer between
structure resulting from the spin–orbit splitting (typically observed in p CdS and TiO2. Ti 2p3/2 and Ti 2p1/2 corresponds to 458.6 and 464.4 eV
and higher orbitals). The peaks at ca. 160 and 161 eV correspond to S (Fig. 8(b)) respectively [45]. Usually, the binding energies of orbital
2p3/2 and S 2p1/2 respectively implying the successful formation of CdS decreases by gaining electrons [46]. Similarly, 403.6 and 410.5 eV
[40]. The S 2p spectra in the composites showed a peak at ca. 168.8 eV corresponds to Cd 3d3/2 and Cd 3d5/2 in Cd 3d spectra (Fig. 8(c)) [47].
that is completely different from the CdS spectra (see Fig. 8(a)). This is The peak shifting towards higher binding energies in the composites
attributed to the bidentate attached SO42− ions with Ti4+ on the sur- indicate the loss of electrons from CdS to TiO2 to form a Z-scheme;
face [41], and the intensity increased with CdS concentration. The ascribed to the difference in Fermi levels [48]. As discussed before, the
SO42− coordinate with TiO2 in the network to give rise to Lewis acidic formation of positive and negative layer at the junction causes the band
sites that capture photo-induced electrons and reduce the recombina- bending due to electrostatic attraction between them. This up-ward
tion [42]. However, the higher number of sites could hamper the band bending forbids the photo-excited electrons to transfer electrons
quantum yield of the process owing to excess trap states favoring re- from CdS conduction band to TiO2. In contrast, TiO2 conduction band
combination. Here, an important aspect to be noted is the oxygen va- electrons migrate to valence band of CdS to recombine with photo-ex-
cancy creation (small amount) resulting from interaction between ti- cited holes [49]. This may result in lowering the probability of re-
tanate and sulfanate in the composites (during the calcination process). combination in same photocatalysts; causing a drastic increase in redox
The identification of oxygen vacancies in the form of Ti3+ will be ex- potentials. Lastly, the electron transfer from CdS to TiO2 is further
plained later. Notably, the peak shifting towards higher energy in S 2p concretized by O 1 s spectra (Fig. 8(d)) where shifting of binding en-
spectra is attributed to the substitutional doping of Ti4+ in the CdS ergies towards lower energy corresponds to electron gain.
lattice causing the charge compensation and giving rise to S-vacancies This can be correlated to a decrease in WF value at lower CdS
[43]. The S-vacancies (Vs) facilitate new energy levels in the forbidden doping to contribute to a better electron transfer. With the further in-
whereupon trapping the electron it becomes Vs− and Vs2− in the con- crease in doping concentration (in R3 and R4), peak shifting was ob-
secutive steps. The transition energies of Vs− and Vs2− are 77 and served towards higher binding energies due to increase in WF as

5
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

Fig. 5. SEM images of (a) CdS, (b) TiO2, (C) R1, (d) R2, (e) R3, and (f) R4.

Fig. 6. Relation between surface dipole moment and work function.

6
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

Fig. 7. Bandgap of catalysts.


Fig. 9. ESR spectra of as-synthesized catalysts: black to navy blue corresponds
(bottom to top) to TiO2, CdS, R1, R2, R3 and R4 respectively.
reflected from UPS measurements (Fig. 3(a, b)). Apart from this, con-
duction and valence band potentials were deduced from equation:
EVBM = ECBM + Eg [50]. The EVBM values for TiO2 (2.85 eV), R1 reduction reactions and support the generation of ROS species. The ROS
(1.99 eV), R2 (1.9 eV), R3 (1.6 eV) and R4 (0.88 eV) were obtained from species (such as superoxide ions and OH radicals) form the basis for the
UPS spectra (Fig. 3a). The conduction band minimum was later calcu- oxidation of organic pollutants. A higher difference between band po-
lated by subtracting VBM values with bandgap. The CBM values for tential and thermodynamic potential drives the photocatalytic reactions
TiO2, R1, R2, R3 and R4 are −0.15, −0.4, −0.44, −0.61, and −1.29 eV at faster rate.
respectively. It can be understood that oxygen reduction reactions are
O2 + e− O2− E = -0.33 V (1)
not possible in TiO2, and R1 owing to thermodynamic constraints (see
Eq. (1)). On the other hand, higher electrical resistance at junction OH− + hVB+ OH* E = 1.99 V (2)
interface blocks the electron movement and thereby leading to charge
H2O + hVB+ OH* + H +
E = 2.34 V (3)
recombination. In R2, the higher reduction potentials drive the Oxygen

Fig. 8. (a) S 2p spectra of CdS, R2, and R4, (b) Ti 2p spectra of TiO2, R2, and R4, (c) Cd 3d spectra of CdS, R3, and R4 and (d) O 1s spectra TiO2, R2, and R4.

7
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

ESR measurements were performed for as-synthesized catalysts and (corresponding to CdS nanoparticles). When these vacancies (here at-
results show that O2− ions formation is more common in the compo- tributed to doping foreign atom) exceed an appropriate concentration,
sites and CdS (due to high reduction potential) as compared to TiO2. the dipoles at the defect states start to interact with neighboring dipoles
However, OH* (radicals) are more prevalent in TiO2 composites (due to to give rise to depolarization (discussed in the earlier sections). The
high oxidation potential EVB = 2.85 eV) as compared to pure CdS. electrostatic energy renormalizes activation and formation energies as
Fig. 9 shows the ESR spectra of TiO2, CdS, and all the reported com- consequence of dipole interaction [60]. This smoothening of electron
posites. The characteristic signals of OH* and O2− ions were recognized charge density over the surface was described as the Smoluchowski
from the standard spectra of OH* (g = 2.0032, B = 3369.1G) [51] and effect and has an important role in the field of surface science. Variation
O2− (g = 2.093 and g = 2.006) [52] respectively. Of all the compo- of local work function values, charge density variation, and irregular
sites, R2 has a clear superoxide ion spectra emphasizing the role of distribution of adsorbates are classical examples of this effect [61]. The
optimal dopant concentration in enhancing the charge transfer and also decrease in work function values (as observed from the UPS spectra)
due to its higher reduction potential. In pure CdS, it is understandable could be attributed to the dipole moment; induced by vacancies (op-
that OH radical formation is limited by its lower oxidation potential. timal concentration). Also, the defect sites have connection to surface
However, in composites, although the lifetime of charge carriers was adsorption and reactivity due to higher coordination at the corrugated
improved, the holes were stuck at the interface; hampering its entry sites. These local properties may assist in binding the surface adsorbates
into TiO2 [49]. Moreover, it can be noticed from results that electrons and increase the reactivity, which was observed in numerous studies
in R2 effectively participate in ORR activity although reaction (3) is associated to oxygen defects [62]. Fig. 11 depicts the charge accumu-
prohibited. In case of R3 and R4, poor electron transfer at interface lation at lower defect sites and depletion at the upper defect sites with
hampered the photocatalytic activity. On the flip side, TiO2 fails to dipole moment Pz.
satisfy the thermodynamic requirement for ORR (Eq. (1)) due to which
the photocatalytic mechanism is completely holes dominated. Never- 3.6. Photocatalytic degradation
theless, in composites, the XPS, ESR and UPS measurements prove that
WF values play a decisive role. This concludes the reaction mechanism To validate the characterization results, the degradation of MB was
of photocatalytic degradation of MB over composites is different from carried out under sunlight with as-synthesized catalysts. The photo-
TiO2. Apart from this, it is noteworthy to understand the absence of degradation results of all the reported catalysts were plotted in
oxygen defects; as CdS cannot penetrate into Ti lattice to replace Fig. 12(a). In the absence of a catalyst (photolysis), there was hardly
oxygen. This makes the overall efficiency of the process highly depen- any change in the MB concentration. From Fig. 12(a) it can be noticed
dent on WF and electron band structures; as there were no modifica- that R2 exhibited the highest MB removal (96%). In this discussion, R3
tions in interface properties attributed to doping. and R4 draw more attention due to sudden underperformance owing to
its higher WF values. Although both TiO2 and CdS are photocatalysts,
3.5. Vacancies as the probable source of surface dipoles the superior performance of former over the latter is attributed to
greater thermodynamic potential differences. In R3 and R4, the en-
The S-vacancies create a difference in charge density on the surface hanced WF values hampered the charge transfer thereby increasing the
to form dipoles [53]. The formation of dipoles assists in charge transfer probability of recombination. The same is reflected in the PL spectra
by favoring the chemisorption of oxygen molecules on the surface. where R3 and R4 have higher intensities owing to band-band re-
Furtherly, these oxygen containing groups transform to reactive oxygen combination. Since CdS is prone to undergo photo corrosion, recycl-
species (ROS). The vacancies form corrugations (due to the missing ability tests were conducted to check the photo-stability of catalysts. In
atom) and tend to spill out electron charges at these sites in order to a Z-scheme mode of charge transfer, the electrons in the conduction
smoothen the abrupt variation of the location of positively charged band of CdS will be deprived of holes (due to its recombination with
nuclei [54]. This can be reasoned by the fact that electrons are less electrons from the TiO2 conduction band) thereby arresting the photo-
localized as compared to nuclei. Electron transporting abilities are corrosion effectively [63]. Fig. 12(b) displays the recyclability de-
naturally enhanced at the interface due to this phenomena. The positive gradation tests for R2 under sunlight. From the results, it can be inferred
shift in binding energies of S 2p spectra (Fig. 10a) reflects the S-va- that there is no substantial change in photocatalytic activity and
cancies [55] and the concentration of S-vacancies increased with CdS therefore the possibility of photo corrosion can be safely ruled out. In
doping. The oxygen vacancies in the composites were identified in R2 this journey of understanding the photo corrosion, an interesting re-
and R4 resulting from sulfate coordination with titanate (as observed velation about the flow of electrons in R2 and R4 was also unveiled and
from S 2p spectra). Usually, Ti3+ is an indicator of oxygen vacancies concretized the discussion in earlier sections. As observed in Fig. 8(c),
[56] in TiO2; as the extra pair of electrons induced by defects reduce the the Cd 3d spectra shifted towards higher binding energy (in R2 and R4)
neighboring Ti4+ ions. Fig. 10b, c indicates the presences of oxygen due to the loss of electrons from CdS to TiO2. However, it should be
vacancies in R2 and R4, which could synergistically complement the S- noted the shift was more pronounced in R2 as compared to R4; sug-
vacancies in all the modifications pertaining to optoelectronics and gesting the hindrance in electron flow from CdS. The higher con-
dissociative adsorption [57]. centration of dipoles in R4 induced the depolarization effect; causing
Moreover, this may also indicate the unlikeliness of Cd penetration the work function value to increase. Moreover, the recyclability results
into Ti lattice; as the oxygen vacancies are formed due to interaction also show R2 as the best performer among the catalysts under con-
between sulfate and titanate in the composite catalysts. Another in- sideration and therefore, the optimum ratios of Cd/Ti also aid in ar-
teresting observation is the decrease in Ti3+ concentration in R4 in resting the photo corrosion. Fig. 13 depicts the role played Z-scheme
comparison to R2. This decrease could be attributed to formation of CdS mode of electron transfer in arresting the recombination.
clusters grown big enough on TiO2 (at higher concentration) and
weakened the interaction between them; due to loose contact [58]. 3.7. Parametric variation
Fig. 10d represents the Raman spectra of TiO2, R2 and R4 where the
chemical shift in the composites were ascribed to OV on anatase. The Furthermore, to understand the influence of process parameters on
characteristic peaks of anatase TiO2 can be identified at ca. 145, 395.9, degradation rate, the concentration of pollutants, catalyst loading and
518.8 and 639.4 cm−1 [59]. It can be understood from Fig. 10d that pH of the solution were systematically varied. Catalyst loading was
with increasing CdS concentration, the peaks at 395.9, 518.8 and varied from 0.3 g/L to 0.9 g/L to determine the optimum loading. From
639.4 cm−1 started diminishing while simultaneously showing the Fig. 14(a) it can be understood that beyond the optimum loading
emergence of new peaks at ca. 260, 425, and 603.5 cm−1 (0.5 g/L), the degradation of MB was limited by aggregation of

8
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

Fig. 10. (a) S 2p spectra of CdS, R2 and R4, Ti3+ identification (induced by OV) and comparison between (b) R2 and (c) R4, and (d) Raman spectra (excitation at
532 nm) of CdS, R2 and R4.

3.8. Defects optimality: To abort undesired thermodynamically favored


back reactions

The higher defect density in the photocatalysts are usually subjected


to thermo-dynamically favored backward reactions. This may lead to
recombination of charge carriers via trap states [16]. Although defects
aid in enhancing the rate of photocatalytic reactions, a higher number
of vacant sites supports charge recombination through multiple trans-
fers from one energy state to another; as indicated from PL spectra of
R3, R4 [65]. Besides this, there are sufficient enough reports that stu-
died a reduction in photocatalytic activity at excess defect density [66].
However, hardly any of them explained how the higher defect density
triggers the undesired thermodynamically favored backward reactions
are triggered. In this section, we demonstrate the step-by-step proce-
dure by accounting all the modifications ascribed to S-vacancies (in the
composites) with an effective comparison between optimum and higher
defect concentration catalysts.
Fig. 11. Schematic representation of surface electric dipole arising from spill- Fig. 15 represents the mechanism of photocatalytic degradation of
out of electrons. MB in R2 and R4 with both photo-excited electrons and holes following
desired/undesired pathways. Cases (a) and (b) discusses the clear
nanoparticles; thereby decreasing the surface area. Additionally, the contrast between catalysts with different defect densities as mentioned.
passage of light could be blocked by the shielding effect [64]. The in- Step (i) shows the higher probability of charge trapping with plenty of
itial concentration of MB was varied to evaluate the uptake capacity of energy states in the forbidden region and found to be more common in
the catalyst (R2). Fig. 14(b) shows the change in % degradation against R4. Though R4 experiences higher trap states, a single catalyst system
change in the initial concentration of MB. The uptake capacities of R2 such as TiO2, CdS lacks an electron cocatalyst at the interface thereby
calculated from the degradation results (0.5 g/L and pH-6) at 20, 30, 40 promoting band to band recombination. This was clearly seen in case
and 50 are 37.2, 46.8, 50.4 and 45 mg/g respectively. Fig. 14(c) dis- (b) (step (v)), where the energy offset for charge transport at interface
plays the pH dependency on overall photocatalytic degradation. was meager; attributed to modification by cocatalyst. On the other
Fig. 14(d) represents the dark experiment that was conducted to ensure hand, thermodynamically undesired backward reactions in case (a)
the adsorption-desorption equilibrium. (step (iii)) may significantly hinder the ORR that produces OH radicals.
But in case (b), due to lower defect density, a higher probability for step

9
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

4. Conclusion

In this work, a series of composite catalysts at different CdS doping


percentages were synthesized. Samples were tested through various
characterizations such as XRD, SEM, XPS, UV-DRS, PL and UPS. Results
showed that an appropriate dopant concentration could dramatically
decrease the work function of the composites. XPS studies reveal the
mode of electron transfer between TiO2 and CdS; confirming the Z-
scheme. Excess concentration of dopant causes a complete over layer on
the surface thereby leading to depolarization. This loss of dipole mo-
ment may cause the electron flow in reverse direction as understood
from UPS results where work function values have drastically in-
creased. Besides, the excess surface defects also have a similar effect to
resist the electron flow at junction interface. Consequentially, band-
band recombination has drastically enhanced at higher Cd doping
concentrations. The composites showed good absorption in the visible
as revealed by bandgap studies. As the mobility of holes was obstructed
by the TiO2 shell, the photocatalytic mechanism was completely
dominated by the oxygen reduction reaction (ORR) in the composites.
Recyclability tests confirms that Z-scheme mode of electron transfer
forms a key in addressing the photo-corrosion in composites. Finally, it
was concluded that excess defect density in the catalysts may have a
detrimental effect on photocatalytic efficiencies.

Declaration of Competing Interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influ-
ence the work reported in this paper.

Fig. 12. Photocatalytic degradation of MB: (a) comparison of catalysts for


overall yield, (b) Recyclability tests. Acknowledgment

This work was supported financially by the Department of Science


(v) can be assumed. In conclusion, the lower defect densities in case (b),
and Technology (DST), Government of India (EMR/2016/003370). We
efficiently addresses the highly probable charge trapping bypassing step
would also like to express our gratitude to Director, BITS-Pilani, K K
(i) and favoring the MB degradation through steps (iv) and (v).
Birla Goa campus for the support to use the institutional infrastructure
for the development of this paper.

Fig. 13. Addressing the photo corrosion by Z-scheme mode of electron transfer.

10
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

Fig. 14. Photocatalytic degradation of MB at varying catalyst loading (a), initial concentration of MB (ppm) (b), and pH (c), (d) dark experiments (adsorption).

Fig. 15. Change in photocatalytic mechanism with defect concentration: (a) R2 and (b) R4.

References 122–138, https://doi.org/10.1002/anie.201705628.


[5] S. Wang, G. Do Lee, S. Lee, E. Yoon, J.H. Warner, Detailed Atomic Reconstruction of
Extended Line Defects in Monolayer MoS2, ACS Nano. 10 (2016) 5419–5430,
[1] K. Takanabe, Photocatalytic Water Splitting: Quantitative Approaches toward https://doi.org/10.1021/acsnano.6b01673.
Photocatalyst by Design, ACS Catal. 7 (2017) 8006–8022, https://doi.org/10.1021/ [6] S. Huang, Y. Long, S. Ruan, Y.J. Zeng, Enhanced Photocatalytic CO2 Reduction in
acscatal.7b02662. Defect-Engineered Z-Scheme WO3- x/g-C3N4 Heterostructures, ACS Omega. 4
[2] D. Maarisetty, S.S. Baral, Defect-induced enhanced dissociative adsorption, optoe- (2019) 15593–15599, https://doi.org/10.1021/acsomega.9b01969.
lectronic properties and interfacial contact in Ce doped TiO2: Solar photocatalytic [7] Y. Zhou, Z. Zhang, Z. Fang, M. Qiu, L. Ling, J. Long, L. Chen, Y. Tong, W. Su,
degradation of Rhodamine B, Ceram. Int. (2019), https://doi.org/10.1016/j. Y. Zhang, J.C.S. Wu, J.M. Basset, X. Wang, G. Yu, Defect engineering of metal-oxide
ceramint.2019.07.251. interface for proximity of photooxidation and photoreduction, Proc. Natl. Acad. Sci.
[3] Y.C. Zhang, N. Afzal, L. Pan, X. Zhang, J.J. Zou, Structure-Activity Relationship of U. S. A. 116 (2019) 10232–10237, https://doi.org/10.1073/pnas.1901631116.
Defective Metal-Based Photocatalysts for Water Splitting: Experimental and [8] R. Shi, Y. Zhao, G.I.N. Waterhouse, S. Zhang, T. Zhang, Defect engineering in
Theoretical Perspectives, Adv. Sci. 6 (2019), https://doi.org/10.1002/advs. photocatalytic nitrogen fixation, ACS Catal. 9 (2019) 9739–9750, https://doi.org/
201900053. 10.1021/acscatal.9b03246.
[4] H. Li, J. Li, Z. Ai, F. Jia, L. Zhang, Oxygen Vacancy-Mediated Photocatalysis of [9] L. Sang, H. Tan, X. Zhang, Y. Wu, C. Ma, C. Burda, Effect of quantum dot deposition
BiOCl: Reactivity, Selectivity, and Perspectives, Angew. Chemie - Int. Ed. 57 (2018) on the interfacial flatband potential, depletion layer in TiO 2 nanotube electrodes,

11
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

and resulting H 2 generation rates, J. Phys. Chem. C. 116 (2012) 18633–18640, Sci. 555 (2019) 801–809, https://doi.org/10.1016/j.jcis.2019.08.036.
https://doi.org/10.1021/jp305388c. [34] B. Sun, Z. Liang, Y. Qian, X. Xu, Y. Han, J. Tian, Sulfur vacancy-rich o-doped 1t-
[10] A.A. Murashkina, P.D. Murzin, A.V. Rudakova, V.K. Ryabchuk, A.V. Emeline, mos2 nanosheets for exceptional photocatalytic nitrogen fixation over cds, ACS
D.W. Bahnemann, Influence of the Dopant Concentration on the Photocatalytic Appl. Mater. Interf.. (2020), https://doi.org/10.1021/acsami.9b20767.
Activity: Al-Doped TiO2, J. Phys. Chem. C. 119 (2015) 24695–24703, https://doi. [35] J.B. Varley, V. Lordi, Electrical properties of point defects in CdS and ZnS, Appl.
org/10.1021/acs.jpcc.5b06252. Phys. Lett. 103 (2013), https://doi.org/10.1063/1.4819492.
[11] X. Cui, W. Xu, Z. Xie, J.A. Dorman, M.T. Gutierrez-Wing, Y. Wang, Effect of dopant [36] A. Sarkar, G.G. Khan, The formation and detection techniques of oxygen vacancies
concentration on visible light driven photocatalytic activity of Sn1-XAgxS2, Dalt. in titanium oxide-based nanostructures, Nanoscale. 11 (2019) 3414–3444, https://
Trans. 45 (2016) 16290–16297, https://doi.org/10.1039/c6dt02812h. doi.org/10.1039/c8nr09666j.
[12] B. Babu, A.N. Kadam, R.V.S.S.N. Ravikumar, C. Byon, Enhanced visible light pho- [37] A. Naldoni, M. Allieta, S. Santangelo, M. Marelli, F. Fabbri, S. Cappelli, C.L. Bianchi,
tocatalytic activity of Cu-doped SnO2quantum dots by solution combustion synth- R. Psaro, V. Dal Santo, Effect of nature and location of defects on bandgap nar-
esis, J. Alloys Compd. 703 (2017) 330–336, https://doi.org/10.1016/j.jallcom. rowing in black TiO 2 nanoparticles, J. Am. Chem. Soc. 134 (2012) 7600–7603,
2017.01.311. https://doi.org/10.1021/ja3012676.
[13] B.R. Tak, S. Dewan, A. Goyal, R. Pathak, V. Gupta, A.K. Kapoor, S. Nagarajan, [38] X. Liu, G. Zhu, X. Wang, X. Yuan, T. Lin, F. Huang, Progress in Black Titania: A New
R. Singh, Point defects induced work function modulation of β-Ga 2 O 3, Appl. Surf. Material for Advanced Photocatalysis, Adv. Energy Mater. (2016), https://doi.org/
Sci. 465 (2019) 973–978, https://doi.org/10.1016/j.apsusc.2018.09.236. 10.1002/aenm.201600452.
[14] L. Matějová, K. Kočí, M. Reli, L. Čapek, A. Hospodková, P. Peikertová, Z. Matěj, [39] R. Jinnouchi, A.V. Akimov, S. Shirai, R. Asahi, O.V. Prezhdo, Upward Shift in
L. Obalová, A. Wach, P. Kuśtrowski, A. Kotarba, Preparation, characterization and Conduction Band of Ta2O5 Due to Surface Dipoles Induced by N-Doping, J. Phys.
photocatalytic properties of cerium doped TiO2: On the effect of Ce loading on the Chem. C. (2015), https://doi.org/10.1021/acs.jpcc.5b06932.
photocatalytic reduction of carbon dioxide, Appl. Catal. B Environ. 152–153 (2014) [40] Y. Wang, X. Xu, W. Lu, Y. Huo, L. Bian, A sulfur vacancy rich CdS based composite
172–183, https://doi.org/10.1016/j.apcatb.2014.01.015. photocatalyst with g-C 3 N 4 as a matrix derived from a Cd-S cluster assembled
[15] M.T. Greiner, L. Chai, M.G. Helander, W.M. Tang, Z.H. Lu, Transition metal oxide supramolecular network for H 2 production and VOC removal, Dalt. Trans. 47
work functions: The influence of cation oxidation state and oxygen vacancies, Adv. (2018) 4219–4227, https://doi.org/10.1039/c7dt04912a.
Funct. Mater. 22 (2012) 4557–4568, https://doi.org/10.1002/adfm.201200615. [41] T. Yamaguchi, Recent progress in solid superacid, Appl. Catal. 61 (1990) 1–25,
[16] T. Hisatomi, K. Domen, Reaction systems for solar hydrogen production via water https://doi.org/10.1016/S0166-9834(00)82131-4.
splitting with particulate semiconductor photocatalysts, Nat. Catal. 2 (2019) [42] H. Li, G. Li, J. Zhu, Y. Wan, Preparation of an active SO42−/TiO2 photocatalyst for
387–399, https://doi.org/10.1038/s41929-019-0242-6. phenol degradation under supercritical conditions, J. Mol. Catal. A Chem. 226
[17] S. Hu, X. Chen, Q. Li, Y. Zhao, W. Mao, Effect of sulfur vacancies on the nitrogen (2005) 93–100, https://doi.org/10.1016/j.molcata.2004.09.028.
photofixation performance of ternary metal sulfide photocatalysts, Catal. Sci. [43] J. Liu, Y. Zhao, J. Liu, S. Wang, Y. Cheng, M. Ji, Y. Zhou, M. Xu, W. Hao, J. Zhang,
Technol. 6 (2016) 5884–5890, https://doi.org/10.1039/c6cy00622a. From Cu2S nanocrystals to Cu doped CdS nanocrystals through cation exchange:
[18] X. Li, Y. Cheng, Q. Wu, J. Xu, Y. Wang, Synergistic effect of the rearranged sulfur controlled synthesis, optical properties and their p-type conductivity research, Sci.
vacancies and sulfur interstitials for 13-fold enhanced photocatalytic H 2 produc- China Mater. 58 (2015) 693–703, https://doi.org/10.1007/s40843-015-0080-z.
tion over defective Zn 2 In 2 S 5 nanosheets, Appl. Catal. B Environ. 240 (2019) [44] F.D. Brandão, G.M. Ribeiro, P.H. Vaz, J.C. González, K. Krambrock, Identification of
270–276, https://doi.org/10.1016/j.apcatb.2018.09.008. rhenium donors and sulfur vacancy acceptors in layered MoS 2 bulk samples, J.
[19] J. Lee, S. Ham, D. Choi, D. Jang, Facile fabrication of porous ZnS nanostructures Appl. Phys. 119 (2016) 235701, , https://doi.org/10.1063/1.4954017.
enhanced photocatalytic performances, Nanoscale. 10 (2018) 14254–14263, [45] B. Bharti, S. Kumar, H.N. Lee, R. Kumar, Formation of oxygen vacancies and Ti3+
https://doi.org/10.1039/c8nr02936a. state in TiO2 thin film and enhanced optical properties by air plasma treatment, Sci.
[20] K. Wu, P. Wu, J. Zhu, C. Liu, X. Dong, J. Wu, G. Meng, K. Xu, J. Hou, Z. Liu, X. Guo, Rep. 6 (2016), https://doi.org/10.1038/srep32355.
Synthesis of hollow core-shell CdS@TiO 2 /Ni 2 P photocatalyst for enhancing [46] F. Xu, J. Zhang, B. Zhu, J. Yu, J. Xu, CuInS 2 sensitized TiO 2 hybrid nanofibers for
hydrogen evolution and degradation of MB, Chem. Eng. J. (2019) 221–230, https:// improved photocatalytic CO 2 reduction, Appl. Catal. B Environ. 230 (2018)
doi.org/10.1016/j.cej.2018.11.211. 194–202, https://doi.org/10.1016/j.apcatb.2018.02.042.
[21] W. Zhou, H. Fu, Defect-mediated electron-hole separation in semiconductor pho- [47] G. Gogoi, S. Keene, A.S. Patra, T.K. Sahu, S. Ardo, M. Qureshi, Hybrid of g-C3N4
tocatalysis, Inorg. Chem. Front. 5 (2018) 1240–1254, https://doi.org/10.1039/ and MoS2 Integrated onto Cd0.5Zn0.5S: Rational Design with Efficient Charge
c8qi00122g. Transfer for Enhanced Photocatalytic Activity, ACS Sustain, Chem. Eng. 6 (2018)
[22] R. Qian, H. Zong, J. Schneider, G. Zhou, T. Zhao, Y. Li, J. Yang, D.W. Bahnemann, 6718–6729, https://doi.org/10.1021/acssuschemeng.8b00512.
J.H. Pan, Charge carrier trapping, recombination and transfer during TiO2 photo- [48] A. Meng, B. Zhu, B. Zhong, L. Zhang, B. Cheng, Direct Z-scheme TiO 2 /CdS hier-
catalysis: An overview, Catal. Today. 335 (2019) 78–90, https://doi.org/10.1016/j. archical photocatalyst for enhanced photocatalytic H 2 -production activity, Appl.
cattod.2018.10.053. Surf. Sci. 422 (2017) 518–527, https://doi.org/10.1016/j.apsusc.2017.06.028.
[23] T. Schultz, N.C. Frey, K. Hantanasirisakul, S. Park, S.J. May, V.B. Shenoy, [49] S. Liu, N. Zhang, Z.R. Tang, Y.J. Xu, Synthesis of one-dimensional CdS@TiO2 core-
Y. Gogotsi, N. Koch, Surface Termination Dependent Work Function and Electronic shell nanocomposites photocatalyst for selective redox: The dual role of TiO2 shell,
Properties of Ti 3 C 2 T x MXene, Chem. Mater. 31 (2019) 6590–6597, https://doi. ACS Appl. Mater. Interf. 4 (2012) 6378–6385, https://doi.org/10.1021/
org/10.1021/acs.chemmater.9b00414. am302074p.
[24] Q. Xu, L. Zhang, J. Yu, S. Wageh, A.A. Al-Ghamdi, M. Jaroniec, Direct Z-scheme [50] D. Maarisetty, J. Komandur, S. Sharma, S.S. Baral, P. Mohapatra, Unravelling the
photocatalysts: Principles, synthesis, and applications, Mater. Today. 21 (2018) rate controlling step in degradation of phenol on a higher potential photocatalyst, J.
1042–1063, https://doi.org/10.1016/j.mattod.2018.04.008. Environ. Chem. Eng. 8 (2020).
[25] W. Jiang, X. Zong, L. An, S. Hua, X. Miao, S. Luan, Y. Wen, F.F. Tao, Z. Sun, [51] D. Maarisetty, S.S. Baral, Synergistic effect of dual electron-cocatalyst modified
Consciously Constructing Heterojunction or Direct Z-Scheme Photocatalysts by photocatalyst and methodical strategy for better charge separation, Appl. Surf. Sci.
Regulating Electron Flow Direction, ACS Catal. 8 (2018) 2209–2217, https://doi. 489 (2019) 930–942, https://doi.org/10.1016/j.apsusc.2019.06.029.
org/10.1021/acscatal.7b04323. [52] E.L.G. Samuel, D.C. Marcano, V. Berka, B.R. Bitner, G. Wu, A. Potter, R.H. Fabian,
[26] M. Reli, N. Ambrožová, M. Šihor, L. Matějová, L. Čapek, L. Obalová, Z. Matěj, R.G. Pautler, T.A. Kent, A.L. Tsai, J.M. Tour, Highly efficient conversion of super-
A. Kotarba, K. Kočí, Novel cerium doped titania catalysts for photocatalytic de- oxide to oxygen using hydrophilic carbon clusters, Proc. Natl. Acad. Sci. U. S. A. 112
composition of ammonia, Appl. Catal. B Environ. 178 (2015) 108–116, https://doi. (2015) 2343–2348, https://doi.org/10.1073/pnas.1417047112.
org/10.1016/j.apcatb.2014.10.021. [53] D. Zhang, M. Yang, S. Dong, Electric-dipole effect of defects on the energy band
[27] O. Gershevitz, C.N. Sukenik, J. Ghabboun, D. Cahen, Molecular monolayer-medi- alignment of rutile and anatase TiO 2, Phys. Chem. Chem. Phys. 17 (2015)
ated control over semiconductor surfaces: Evidence for molecular depolarization of 29079–29084, https://doi.org/10.1039/c5cp04495b.
silane monolayers on Si/SiOx, J. Am. Chem. Soc. 125 (2003) 4730–4731, https:// [54] O.P. Polyakov, M. Corbetta, O.V. Stepanyuk, H. Oka, A.M. Saletsky, D. Sander,
doi.org/10.1021/ja029529h. V.S. Stepanyuk, J. Kirschner, Spin-dependent Smoluchowski effect, Phys. Rev. B -
[28] D. Cornil, Y. Olivier, V. Geskin, J. Cornil, Depolarization effects in self-assembled Condens. Matter Mater. Phys. 86 (2012) 2–5, https://doi.org/10.1103/PhysRevB.
monolayers: A quantum-chemical insight, Adv. Funct. Mater. 17 (2007) 86.235409.
1143–1148, https://doi.org/10.1002/adfm.200601116. [55] G. Wang, B. Huang, Z. Li, Z. Lou, Z. Wang, Y. Dai, M.-H. Whangbo, Synthesis and
[29] C.S. Kwok, P.D. Mourad, L.A. Crum, B.D. Ratner, Surface modification of polymers characterization of ZnS with controlled amount of S vacancies for photocatalytic H2
with self-assembled molecular structures: Multitechnique surface characterization, production under visible light, Sci. Rep. 5 (2015) 8544, https://doi.org/10.1038/
Biomacromolecules. 1 (2000) 139–148, https://doi.org/10.1021/bm000292w. srep08544.
[30] G. Zhu, L. Pan, T. Xu, Z. Sun, CdS/CdSe-cosensitized TiO2photoanode for quantum- [56] B. Bharti, S. Kumar, H.N. Lee, R. Kumar, Formation of oxygen vacancies and
dot-sensitized solar cells by a microwave-assisted chemical bath deposition method, Ti3+state in TiO2thin film and enhanced optical properties by air plasma treat-
ACS Appl. Mater. Interf.. 3 (2011) 3146–3151, https://doi.org/10.1021/ ment, Sci. Rep. 6 (2016) 1–12, https://doi.org/10.1038/srep32355.
am200648b. [57] D. Maarisetty, S. Mahanta, A.K. Sahoo, P. Mohapatra, S.S. Baral, Steering the charge
[31] N. Zhang, Y. Zhang, X. Pan, M.Q. Yang, Y.J. Xu, Constructing ternary CdS-graphene- kinetics in dual-functional photocatalysis by surface dipole moments and band edge
TiO 2 hybrids on the flatland of graphene oxide with enhanced visible-light pho- modulation: A defect study in TiO2-rGO-ZnS composite, ACS Appl, Mater. Interf.
toactivity for selective transformation, J. Phys. Chem. C. 116 (2012) 18023–18031, (2020), https://doi.org/10.1021/acsami.9b22418 acsami.9b22418.
https://doi.org/10.1021/jp303503c. [58] M.J. Muñoz-Batista, M.N. Gómez-Cerezo, A. Kubacka, D. Tudela, M. Fernández-
[32] H. Ge, F. Xu, B. Cheng, J. Yu, W. Ho, S-scheme heterojunction TiO2/CdS nano- García, Role of interface contact in CeO2-TiO2 photocatalytic composite materials,
composite nanofiber as H2-production photocatalyst, ChemCatChem. (2019) 1–10, ACS Catal. 4 (2014) 63–72, https://doi.org/10.1021/cs400878b.
https://doi.org/10.1002/cctc.201901486. [59] Q. Zhang, L. Wang, J. Feng, H. Xu, W. Yan, Enhanced photoelectrochemical per-
[33] A. Mahmood, J.-W. Park, TiO2/CdS nanocomposite stabilized on a magnetic-cored formance by synthesizing CdS decorated reduced TiO 2 nanotube arrays, Phys.
dendrimer for enhanced photocatalytic activity and reusability, J. Colloid Interf. Chem. Chem. Phys. 16 (2014) 23431–23439, https://doi.org/10.1039/

12
D. Maarisetty, et al. Applied Surface Science 536 (2021) 147843

C4CP02967D. Degrading Water Dyes and Antibiotics, ACS Sustain. Chem. Eng. 5 (2017)
[60] M. Giesen, An atomistic view on fundamental transport processes on metal surfaces, 6958–6968, https://doi.org/10.1021/acssuschemeng.7b01157.
AIP Conf. Proc. 916 (2007) 115–135, https://doi.org/10.1063/1.2751912. [64] S.P. Kamble, S.B. Sawant, V.G. Pangarkar, Batch and Continuous Photocatalytic
[61] H.J. Lee, A.C. Jamison, T.R. Lee, Surface Dipoles: A Growing Body of Evidence Degradation of Benzenesulfonic Acid Using Concentrated Solar Radiation, Ind. Eng.
Supports Their Impact and Importance, Acc. Chem. Res. 48 (2015) 3007–3015, Chem. Res. 42 (2003) 6705–6713, https://doi.org/10.1021/ie030493r.
https://doi.org/10.1021/acs.accounts.5b00307. [65] R. Yew, S.K. Karuturi, J. Liu, H.H. Tan, Y. Wu, C. Jagadish, Exploiting defects in TiO
[62] D.N. Pei, L. Gong, A.Y. Zhang, X. Zhang, J.J. Chen, Y. Mu, H.Q. Yu, Defective ti- 2 inverse opal for enhanced photoelectrochemical water splitting, Opt. Exp. 27
tanium dioxide single crystals exposed by high-energy 001 facets for efficient (2019) 761, https://doi.org/10.1364/oe.27.000761.
oxygen reduction, Nat. Commun. 6 (2015) 1–10, https://doi.org/10.1038/ [66] M. Kong, Y. Li, X. Chen, T. Tian, P. Fang, F. Zheng, X. Zhao, Tuning the Relative
ncomms9696. Concentration Ratio of Bulk Defects to Surface Defects in TiO 2 Nanocrystals Leads
[63] Q. Li, Z. Guan, D. Wu, X. Zhao, S. Bao, B. Tian, J. Zhang, Z-Scheme BiOCl-Au-CdS to High Photocatalytic Efficiency, J. Am. Chem. Soc. 133 (2011) 16414–16417,
Heterostructure with Enhanced Sunlight-Driven Photocatalytic Activity in https://doi.org/10.1021/ja207826q.

13

You might also like