You are on page 1of 48

Journal Pre-proof

Effect of synthesis method parameters on properties


and photoelectrocatalytic activity under solar
irradiation of TiO 2 nanotubes decorated with CdS
quantum dots

Aleksandra Pieczyńska, Paweł Mazierski,


Wojciech Lisowski, Tomasz Klimczuk, Adriana
Zaleska-Medynska, Ewa Siedlecka

PII: S2213-3437(20)31165-9
DOI: https://doi.org/10.1016/j.jece.2020.104816
Reference: JECE104816

To appear in: Journal of Environmental Chemical Engineering


Received date: 3 October 2020
Revised date: 17 November 2020
Accepted date: 19 November 2020
Please cite this article as: Aleksandra Pieczyńska, Paweł Mazierski, Wojciech
Lisowski, Tomasz Klimczuk, Adriana Zaleska-Medynska and Ewa Siedlecka,
Effect of synthesis method parameters on properties and photoelectrocatalytic
activity under solar irradiation of TiO 2 nanotubes decorated with CdS quantum
d o t s , Journal of Environmental Chemical Engineering, (2020)
doi:https://doi.org/10.1016/j.jece.2020.104816
This is a PDF file of an article that has undergone enhancements after acceptance,
such as the addition of a cover page and metadata, and formatting for readability,
but it is not yet the definitive version of record. This version will undergo
additional copyediting, typesetting and review before it is published in its final
form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2020 Published by Elsevier.
Effect of synthesis method parameters on properties and photoelectrocatalytic

activity under solar irradiation of TiO2 nanotubes decorated with CdS quantum

dots

Aleksandra Pieczyńskaa*, Paweł Mazierskia, Wojciech Lisowskib, Tomasz Klimczukc, Adriana

Zaleska-Medynska a, Ewa Siedlecka a

f
a
Department of Environmental Technology, Faculty of Chemistry, University of Gdansk, 80-308

oo
Gdansk, Poland

pr
b
Institute of Physical Chemistry, Polish Academy of Science, Kasprzaka 44/52, 01-244 Warsaw,

Poland e-
c
Department of Solid State Physics, Gdansk University of Technology, 80-233 Gdansk, Poland
Pr
al

*Corresponding Author
n

Email : aleksandra.pieczynska@ug.edu.pl
ur
Jo

Abstract

The growing research interest on photoelectrocatalysis has encouraged the search for new

materials with high activity and the development of methods for their synthesis. The successive

ionic layer adsorption and reaction (SILAR) method is an effective way to synthesize materials

with photoelectrocatalytic (PEC) properties that are active under visible radiation. Therefore,

studies on the impact of the parameters of the SILAR method on the properties and PEC activity

of TiO2 nanotubes sensitized with CdS quantum dots (QDs) were conducted. PEC activity was

1
determined based on the efficiency of ifosfamide (IF) degradation, which is currently one of the

most commonly detectable anticancer drugs in aquatic environment. The highest IF PEC

degradation rate (0.0184 min-1) and the highest TOC removal (48%) was achieved with the CdS-

Ti/TiO2_NO3-_0.1M_3c_60s photoelectrode. Among the SILAR parameters, the highest impact

on the PEC IF degradation efficiency had the concentration of the Cd precursor, as well as the

number and time of the cycles. Our studies showed that CdS-Ti/TiO2 nanocomposites exhibited

f
more than twice higher photocatalytic and PEC IF degradation efficiency than pristine Ti/TiO2

oo
under UV-Vis irradiation. In addition, the PEC mechanism of IF degradation using pristine

pr
Ti/TiO2 and CdS-Ti/TiO2 was compared, from which it was concluded that the higher PEC

activity of the nanocomposites is due to the greater contribution of h+ and the generation of other
e-
oxidants caused by the presence of CdS QDs. Based on the identified degradation products, a
Pr
pathway for PEC IF degradation was also proposed.
al

Kaywords: photoelectrocatalysis, TiO2 nanotubes, CdS quantum dots, SILAR method,


n

ifosfamide
ur

1. Introduction
Jo

Highly ordered TiO2 nanotubes (NTs) have been widely studied for their application in various

fields, such as in the photocatalytic (PC) and photoelectrocatalytic (PEC) degradation of organic

pollutants, hydrogen production, photoconversion of CO2, and solar cells [1], owing to their

unique properties compared to other TiO2 nanoparticles, including a large surface area to volume

ratio and a relatively low recombination rate of photogenerated electron-hole (e-/h+) pairs.

However, their poor visible light absorption and recombination rate of photogenerated e-/h+ pairs

2
limit their practical application. There are different methods to extend their PC activity into the

visible light region. Some of these involve doping with other elements like N [2,3], B [4], and C

[5], surface modification with metals like Pd [6] and Pt [7], and using the quantum dots (QDs) of

narrow band semiconductors like CdS [8,9], Bi2S3 [10], and PbS [11]. Among the semiconductor

QDs, CdS, with an Eg of 2.4 eV, is one of the most advantageous in practical applications

because of its relatively high absorption coefficient in the visible region of the spectrum, as well

f
as its low cost and simple preparation method [8,9]. The sensitization of TiO2 with CdS QDs

oo
produces nanocomposites which exhibit visible and ultraviolet (UV) light absorption and an

pr
improved separation of photogenerated charges [12].

e-
It has been reported that QDs can be immobilized onto the TiO2 NT surface through

electrochemical deposition, sequential bath deposition, and successive ionic layer adsorption and
Pr
reaction (SILAR). SILAR is a cost-effective and straightforward method for the deposition of

thin films [9]. More importantly, in this method, it is possible to control the size and amount of
al

the particles deposited. The PC activity of nanocomposites with QDs depends on the size,
n

amount, and distribution of the QDs. The parameters of the SILAR method which control the
ur

properties of the deposited QDs are the number of cycles, time of dipping in the precursor
Jo

solutions, and the type and concentration of the precursors [13]. The increase in the number of

SILAR cycles leads to the increase in the amount and the size of the deposited QDs. However,

high numbers of SILAR cycles can result in the aggregation of QDs, which decreases the PC

activity. Therefore, obtaining the optimal number of SILAR cycles is important in the fabrication

of new materials. The size of QDs also depends on the time of dipping in precursor solutions.

Increasing the time for adsorption gradually increases the particle size of the QDs. The type and

concentration of the anion and cation precursors also influence the properties of the QDs. In the

3
case of a CdS QD, the S2- precursor usually comes from a Na2S solution, and the Cd2+ precursor

may be derived from a CdSO4, Cd(NO3)2, or CdCl2 solution. The higher the concentration of the

precursor is, the greater the number and size of QDs particles.

Thus far, TiO2 decorated with CdS QDs is mainly used in solar cells [10,13,14] and as

photocatalysts [9,15,16]. It was also reported that in the PC process, CdS-Ti/TiO2

nanocomposites exhibited a higher activity of degradation of organic compounds than pure TiO2

f
oo
under visible light. However, there is limited information about the application of TiO2 NTs

decorated with CdS QDs as a photoelectrode in PEC degradation. To the best of our knowledge,

pr
there is still a lack of comprehensive investigation on the preparation of CdS-Ti/TiO2, and it
e-
remains unclear how their morphology affects the PEC efficiency, as well the PEC mechanism.
Pr
Hence, in the present study, the effect of the parameters of the SILAR method (type and

concentration of Cd2+ precursor, number of cycles, and dipping time) employed in the deposition
al

of CdS QDs on the surface of Ti/TiO2 NTs on the properties and the PEC activity of CdS-

Ti/TiO2 photoelectrodes was extensively investigated. The structure and morphology of CdS-
n

Ti/TiO2 were investigated using scanning electron microscopy (SEM), transmission electron
ur

microscopy (TEM), X-ray diffraction (XRD) analysis, X-ray photoelectron spectroscopy (XPS),
Jo

ultraviolet-visible (UV-Vis) spectroscopy, and photoelectrochemical analysis.

The PEC degradation activity of the prepared electrodes was determined using an ifosfamide

(IF) solution. IF is one of the most widely used anticancer drugs and has already been detected in

environmental and wastewater samples at the level of ng L-1 [17]. Although, the concentration in

environmental so far is not high, taking into account the increasing amount of consumed

anticancer drugs, it will increase. While IF shows no toxicity towards bacteria and algae, it may

be cytotoxic for other organisms [18]. Therefore, effective IF removal from the wastewater

4
would reduce its distribution in the environment and the risks associated with it. To the best of

our knowledge, there have been no studies on the PEC oxidation of this cytostatic drug.

Moreover, the mechanism of PEC degradation based on the participation of individual oxidants

in IF degradation determined by reactions with scavengers and the pathway of IF PEC

decomposition determined by identified degradation products were proposed.

2. Materials and methods

f
oo
2.1. Chemicals and materials

Ti foils (0.127 mm thickness, 99.7% purity) were purchased from Sigma-Aldrich.

pr
Isopropanol, acetone, and methanol with practical grade were sourced from P.P.H.
e-
STANLAB. Ethylene glycol (EG) (99.0%), CdCl2·2.5H2O, CdSO4·H2O, Na2S·9H2O, and
Pr
Cd(NO3)2·H2O was obtained with analytical grade from CHEMPUR. Analytical grade

ammonium fluoride was purchased from Acros Organics. The ifosfamide (99%) standard and
al

terephthalic acid (98%) were obtained from Sigma-Aldrich (Steinheim, Germany).

Analytical grade sodium sulfate, ascorbic acid, formic acid, and methanol were purchased
n

from POCh S.A. (Gliwice, Poland). High-performance liquid chromatography (HPLC) grade
ur

acetonitrile from POCh S.A. Poland was used for the chromatographic measurements.
Jo

2.2. Preparation of CdS-Ti/TiO2 photoelectrodes

TiO2 NTs were prepared via the electrochemical anodization of Ti foil, according to our

previous work [19]. In brief, Ti foils (2 x 5 cm) were cleaned successively using acetone,

isopropanol, methanol, and deionized water for 10 min before use and then dried under an air

stream. Anodization was carried out using a three-electrode electrochemical set-up, where a

Ti foil was the working electrode, Pt mesh was the counter electrode, and Ag/AgCl was the

5
reference electrode, in an electrolyte composed of EG, ammonium fluoride (0.09 M), and

H2O (2 vol.%) at 30 V set using a programmable power supply (MCP M10-QS1005). The

obtained samples were flushed with deionized water, sonicated in water for 5 min, dried in

air (80 °C, 24 h), and calcined at 450 °C for 1 h with heating rate 2 °C min-1.

CdS QDs were deposited onto TiO2 NTs using the SILAR method. The as-prepared TiO2

NTs were immersed in a solution containing the Cd precursor for a specified time, and then

f
oo
the samples were rinsed with pure ethanol and dried. Subsequently, the samples were

immersed into a 0.05 M Na2S·9H2O methanol solution for a specified time, rinsed with

pr
methanol, and then dried. This procedure was repeated for a specified number of times, and
e-
finally, the samples were dried at 80 °C. The preparation conditions are summarized in Table

1.
Pr
2.3. Characterization of the CdS-Ti/TiO2 photoelectrodes
al

The morphology of the prepared samples was characterized using field emission SEM (JSM-
n

7610F, JEOL) and high-resolution TEM (Hitachi H-800) at an accelerating voltage up to 150
ur

kV. For these tests, 0.5 cm x 0.5 cm samples were used. The samples for TEM analysis were

prepared by dry transferring to a carbon support mesh. The length of TiO2 nanotubes could
Jo

be measured after releasing them from the titanium foil by shear forces acting during the

cutting of the foil with a scalpel. The developed surface area (S) was calculated using the

following equation (Eq. 1):

8πhR 2
S= S Eq. 1
3(4R 2 −2R 1 −y)2 geo

where Sgeo is the geometric surface area (8 cm2), R1 and R2 are the internal and external radii,

respectively, y is the distance between the tubes (0 nm), and h is the length of tubes [20].

6
Powder XRD (pXRD) was used to determine the crystal structure and calculate the lattice

parameters of TiO2-anatase. The experiment was performed at room temperature on a D2

Phaser (Brücker) diffractometer (U = 30 kV, I = 10 mA), using Cu Kα radiation (λ = 1.54056

Å) and a LynxEye XE-T detector. Data were collected at a scanning angle 2θ = 10−90° and a

step size 0.01°. The lattice parameters for TiO2-anatase were estimated using the LeBail

profile method and the High Score Plus software.

f
oo
XPS measurements were performed using a PHI 5000 VersaProbe (ULVAC-PHI)

spectrometer with Al Kα radiation (h = 1486.6 eV). High-resolution XPS spectra were

pr
collected with the hemispherical analyzer at the pass energy of 23.5 eV and the energy step
e-
size of 0.1 eV. The binding energy (BE) scale of all detected spectra was referenced by

setting the BE of the aliphatic carbon peak (C-C) signal to 284.8 eV.
Pr
The UV-Vis absorbance spectra of the CdS-Ti/TiO2 photoelectrodes were obtained using a
al

Shimadzu UV-Vis Spectrophotometer (UV 2600) in the range of 300−800 nm, with a

scanning speed of 250 nm min-1 .


n
ur

The electrochemical and photoelectrochemical properties of the as-prepared photoelectrodes

were investigated using an AutoLab PGSTAT 204 potentiostat-galvanostat (Methrom


Jo

Autolab) in a three-electrode system, where the CdS-Ti/TiO2 photoelectrodes, Ag/AgCl/0.1

M KCl, and Pt mesh were used as the working, reference, and counter electrodes,

respectively. The electrolyte composed of 0.1 M sodium sulfate solution was purged with

argon for 1 h. The chronoamperometry measurements under UV-Vis were performed using a

150 W xenon lamp (Hamamatsu Photonics K.K., Model E7536) equipped with a water

infrared cut-off filter. The irradiation intensity was measured using an optical power meter

7
(Hamamatsu, C9536-01) and adjusted to 100 mW cm-2. Cyclic voltammetry analysis was

performed in the dark by applying a potential from -1 to 2 V.

The amount of hydroxyl radicals (∙OH) generated in the PEC application of the CdS-Ti/TiO2

photoelectrodes were determined using a fluorescence technique according to the procedure

presented by Mazierski et al. [19].

2.4. Photoelectrocatalytic oxidation of IF

f
oo
The PEC degradation of IF was performed in a photoreactor with a three-electrode

configuration, with one of the CdS-Ti/TiO2 electrodes (surface area 4 cm2) as the anode,

pr
stainless steel as the cathode, and Ag/AgCl/0.3 M KCl as the reference electrode. To provide
e-
a constant potential at 1 V an Aim TTi PL303 power supply (Huntingdon, England) was
Pr
employed. A Suntest CPS+ solar simulator (Atlas Material Testing Technology LLC)

equipped with a xenon lamp was used as the UV-Vis irradiation source with an intensity of
al

550 W m-2. In each process, 80 mL of 20 mg L-1 IF solution in 0.1 M sodium sulfate solution
n

(the supporting electrolyte) were used.


ur

Additionally, the photolysis (without photoelectrode and applied potential), electrolysis


Jo

(without irradiation), and photocatalysis (without applied potential) of IF with CdS-

Ti/TiO2_NO3-_0.1M_3c_60s and Ti/TiO2 photoelectrodes were investigated.

2.5. Analytical methods

The concentrations of IF during all processes were determined using HPLC with a UV

detector (Perkin Elmer, Series 200), and the degradation products were identified using an

Agilent 1200 Series LC system (Agilent Technologies, Inc., Santa Clara, USA) coupled to an

8
HCT Ultra ion trap mass spectrometer (Brucker Daltonics, Bremen, Germany), according to

the method of Siedlecka et al. [21].

Total organic carbon (TOC) were measured using a TOC analyzer (TOC-L, TNM-L,

Shimadzu, GmbH, Germany) equipped with an autosampler ASI-L. The mineralization

current efficiency (MCE) for IF based on the TOC data, according to Siedlecka et al. [21],

was estimated from the Eq. 2:

f
𝑛 𝐹 𝑉𝑠 (∆𝑇𝑂𝐶)𝑒𝑥𝑝
𝑀𝐶𝐸 % = 100 Eq. 2

oo
4.32 10 7 𝑚 𝐼 𝑡

where n is the number of electrons consumed per one molecule of IF, assuming total

pr
mineralization (according to Reaction (1)), F is the Faraday constant (96,487 C mol-1), Vs

(dm3) is the solution volume, (ΔTOC)exp (mg L-1) is the experimental difference between the
e-
TOC before and after the PEC process, 4.32 x 107 is a factor to homogenize units (3,600 s h-1
Pr
x 12,000 mg mol-1), m is the number of carbon atoms of IF (mIF = 7), I (A) is the current, and

t (h) is time elapsed in the PEC process.


al

𝐶7 𝐻15 𝑁2 𝑂2 𝐶𝑙2 𝑃 + 16𝐻2 𝑂 → 7𝐶𝑂2 + 2𝑁𝐻4+ + 2𝐶𝑙− + 𝑃𝑂43− + 39𝐻 + + 36𝑒 − (1)
n

The amount of electrical energy (EEO, kWh m-3 order-1) consumed for the PEC and PC
ur

degradation of IF, with initial concentration C0 and concentration after treatment Ct, were
Jo

calculated according to Eq. 3 [22]:

𝑃𝐿𝑎𝑚𝑝 +𝑃𝐶𝑒𝑙𝑙 𝑡
𝐸𝐸𝑂 = 𝐶
Eq. 3
𝑉𝑠 log 𝐶0
𝑡

where PLamp (kW) is the power of the lamp used for irradiation, PCell (kW) is the power

electrochemical cell, t (h) is the time elapsed in the process, and Vs (m-3) is the volume of

treated solution.

3. Results

9
3.1. Characteristics of CdS-Ti/TiO2 photoelectrodes

3.1.1. Scanning electron microscopy and transmission electron microscopy

Fig. 1 shows the surface morphologies of the obtained photoelectrodes. As seen in the SEM

images, the TiO2 NT layers were top-end-open, vertically oriented, and uniform, and these

characteristics were observed on almost all Ti sheets. The sample CdS-Ti/TiO2_Cl-

_0.1M_3c_60s showed a different morphology. The inner walls and the top layer of the

f
nanotube arrays were covered by a CdS layer formed by aggregated QDs rather than by CdS

oo
QDs homogeneously distributed over the surface of the nanotubes. This is confirmed by the

pr
XPS analysis, the Cd precursor in the form of CdCl2·2.5H2O allowed deposition on the

nanotubes surface the largest amount (21.1 at.%) of CdS (among the examined Cd
e-
precursors), which contributed to formation of QDs in the form of big aggregates. For the
Pr
remaining samples, the deposition of the CdS QDs did not influence the morphology of the

NT layers (Fig. 1). Presumably, this is because the CdS layer is decorated with QDs with a
al

small size and amount (explained further in XRD and TEM analysis), relatively thin, and
n

placed inside and outside of the tubes (in other words, along the vertical orientation of the
ur

tubes) rather than at the top layer. The TiO2 NT layers had an average length of 1.7 μm and a
Jo

tube diameter of 65 nm. Thus, the surface area all of the obtained photoelectrodes reached

the value of 549 cm2.

Based on the TEM analysis (Fig. 2.), it was confirmed that the CdS deposited on the TiO2

NTs was in the form of QDs with dimensions not exceeding 6 nm (Table A2 (Supplementary

data)). The CdS QDs were distributed both inside and outside of the NTs. It was found that

the sizes of the QDs on the photoanode depend on the parameters of the SILAR method.

When CdCl2, CdSO4, and Cd(NO3)2 were used as the Cd precursor, the diameters of the CdS

10
QDs were 5.2, 3.6, and 3.1 nm, respectively. As the precursor concentration, number of

SILAR cycles, and dipping time were increased, the size of the QDs also increased.

Generally, the increase in the size of the QDs observed under TEM was consistent with the

increase in the amount of CdS determined in the XPS analysis (see details in Section 3.1.3).

3.1.2. X-ray diffraction (XRD)

The obtained pXRD patterns are presented in Fig. 3. As expected, only two phases were clearly

f
oo
visible: the Ti metal and the TiO2-anatase form. The red and black vertical bars below the

patterns represent the Bragg positions for TiO2 and Ti, respectively. The blue line through the

pr
data points (open circles) is the LeBail fit in which the two phases were included. The estimated
e-
lattice parameters for TiO2 are summarized in Table A1 (Supplementary data). The presence of

CdS QDs was indicated by the peak at around 27°. The hump was rather wide, suggesting the
Pr
small size of the QDs; however, the qualitative analysis to confirm the crystallite size cannot be

carried out. In order to compare the relative signal of CdS for all tested samples, pXRD patterns
al

for 22.5° < 2 < 35° are shown in the right panels of the figures. The relative intensity scale was
n

changed by the same factor.


ur

Fig. 3a. shows the pXRD patterns for the samples prepared using the three different Cd precursor
Jo

solutions: CdCl2, CdSO4, and Cd(NO3)2. As expected, there was no peak near 27° as observed

for a reference sample Ti/TiO2. CdS QDs were visible in the panels of CdS-Ti/TiO2_Cl-

_0.1M_3c_60s and Ti/TiO2_NO3-_0.1M_3c_60s, while the peak was less pronounced for the

samples prepared with CdSO4. Fig. 3b shows the pXRD patterns for the samples prepared with

different concentrations of Cd(NO3)2 solution. An intense and broad reflection near 27° was seen

in the pattern for the sample prepared with the highest concentration (0.2 M) of Cd(NO3)2. This

suggests that the amount of CdS QD increases with the Cd(NO3)2 concentration. Fig. 3c shows

11
how the number of the SILAR cycles influences the presence of CdS QDs. As expected, the

more SILAR cycles carried out, the higher the intensity of the broad reflection associated with

CdS. As observed in Fig. 3d, the relative amount of CdS QDs in the samples prepared by dipping

in the precursor Cd(NO3)2 solution for 0.5 min and 1 min was almost the same. Dipping a sample

for 2 minutes causes a substantial increase in the CdS reflection. In summary, long and precise

pXRD patterns allow us to study the effect of several parameters to the development of CdS

f
QDs. The most efficient precursors are CdCl2 and Cd(NO3)2, while the weakest precursor is

oo
CdSO4. Increasing the precursor concentration, dipping time, and number of SILAR cycles

pr
causes the increase in the relative amount of CdS QDs, as indicated by the increase in the

intensity of the broad peak near 27° associated to the strongest (111) reflection of CdS.
e-
3.1.3. X-ray photoelectron spectra (XPS)
Pr
The elemental surface composition of Ti/TiO2 and the CdS-Ti/TiO2 samples, as evaluated using
al

XPS, is presented in Table A2 (Supplementary data). The high resolution Ti 2p, O 1s, Cd 3d, and

S 2p XPS spectra were recorded to detect Ti, O, Cd, and S, respectively. A sample set of these
n

spectra, collected for the CdS-Ti/TiO2_NO3-_0.2M_3c_60 sample, is shown in Fig. A1


ur

(Supplementary data). The Cd 3d and S 2p spectra confirmed the effective deposition of CdS
Jo

QDs onto the TiO2 NTs [23–25]. The main Cd 3d3/2 signal located at 405.2 eV, can also be

assigned to CdCO3 and CdSO4 [23]. The presence of the last state was also confirmed by the S

2p spectrum (Fig. A1 (Supplementary data)). The main S 2p3/2 signal at 161.4 eV, characteristic

for CdS [23–25], was accompanied by signals located in the BE range of 166−169 eV, which are

assigned to the SOx species (sulfite and sulfate) [23]. The deconvolution of the S 2p spectrum

allows the CdS fraction to separate, and its contribution to the surface chemical composition of

CdS-Ti/TiO2 samples can then be evaluated (see CdS data in Table A2 (Supplementary data)).

12
Moreover, the S 2p spectrum revealed the relative amount of CdS is determined by the

parameters of the SILAR method (Fig. 4). The influence of the type of the Cd precursor (based

on Cl-, SO42-, and NO3- ions) on the amount of CdS is shown in Fig. 4a. The CdS-Ti/TiO2_Cl-

_0.1M_3c_60s sample had the largest amount of CdS compared with the CdS-Ti/TiO2_ SO42-

_0.1M_3c_60s and CdS-Ti/TiO2_NO3-_0.1M_3c_60s samples (compare also the CdS data in

Table A2 (Supplementary data)). The effect of precursor concentration for the samples based on

f
NO3- ions is presented in Fig. 4b. It was observed that the concentration of detected CdS

oo
fractions increased relative to their nominal values. The effect of number of SILAR cycles and

pr
dipping time is shown in Fig. 4c and Fig. 4d, respectively. Evidently, both the increase in the

number of SILAR cycles (from 2 to 5) and dipping time of the CdS-Ti/TiO2 photoanodes (from
e-
30 s to 120 s) led to the relative increase in the amount of CdS (Table A2 (Supplementary data)).
Pr
3.1.4. UV-Vis absorbance
al

In Fig. 5 the UV-Vis spectra of Ti/TiO2 and the CdS-Ti/TiO2 photoanodes with different CdS

deposition parameters of the SILAR method are presented. The Ti/TiO2 absorption spectra had
n

two peaks (Fig. 5a); one below 400 nm, which is characteristic for anatase TiO2, and the other in
ur

the visible light region, probably associated with the trapped electrons at the Ti3+ center [19]. In
Jo

all photoanodes modified with CdS, a strong absorption was observed in the visible region from

380 nm to 550 nm. This is in line with previous studies [9,16,26] which indicate extension

absorbing radiation in the visible range of the TiO2 decorated CdS. Moreover, photoanodes

absorption intensity strongly depends on the parameters of the CdS deposition method. As shown

in Fig. 5a, among the electrodes prepared with different Cd precursors, the sample CdS-

Ti/TiO2_NO3-_0.1M_3c_60s prepared using Cd(NO3)2 exhibited the strongest absorbance of

visible light. The lowest absorbance of visible light was observed with Ti/TiO2_Cl-

13
_0.1M_3c_60s photoelectrode. This could be related to the agglomeration of CdS QDs into large

aggregates on TiO2 surface which does not expand in increasing radiation absorption [14]. No

correlation was observed when the absorption of samples prepared with different concentrations

of Cd precursors (Fig. 5b) were compared. The photoelectrode prepared with 0.1 M of Cd(NO3)2

exhibited the strongest absorption. In the case of the Ti/TiO2_NO3-_0.05M_3c_60s

photoelectrode the lowest absorption of visible radiation was observed as a result in the lowest

f
amount of CdS QDs. The number of SILAR cycles and dipping time had a minor effect on the

oo
absorption spectra of the photoanodes (Fig. 5c and 5d.). Our investigations showed that more the

pr
SILAR cycles and in the same time higher amount of CdS lowers the absorption intensity, which

is in contradiction to Lan et al. [14]. On the other hand, longer dipping times resulted in
e-
increased CdS concentration on Ti/TiO2 shows higher absorption of visible radiation. In
Pr
conclusion, by sensitizing Ti/TiO2 with CdS QDs, its photo-response was extended to the visible

light region. The type of precursor and their concentration had the greatest impact on the shape
al

of the absorption spectra of the photoelectrodes prepared using the SILAR method. However,
n

there was no correlation between the amount of CdS QDs and the intensity of visible radiation
ur

absorption. It is worth noting that, in addition to the quantity of QDs, their size also affects the
Jo

absorption of irradiation [9].

3.1.5. Photoelectrochemical analyses

The PEC property is an important parameter that shows the efficiency of photogenerated charge

separation (e- and h+). The higher the photocurrent is, the better the charge separation and

activity [3,27]. The photocurrents obtained at 1 V and the UV-Vis irradiation of Ti/TiO2 and

CdS-Ti/TiO2 (Fig. 6) demonstrated the photoelectrochemical repeatability and stability of the

photoelectrodes. In addition, higher photocurrent generation was observed in all CdS-Ti/TiO2

14
photoelectrodes compared to pristine Ti/TiO2 (Fig. 6a), which is the result of the formation of

more charges on both the TiO2 and CdS semiconductor. Comparing the efficiency of

photocurrent generation of the photoelectrodes prepared using different Cd precursors (Fig. 6a),

the highest photocurrent value of 3 mA cm-2 was obtained for the CdS-Ti/TiO2_NO3-

_0.1M_3c_60s photoelectrode, followed by CdS-Ti/TiO2_Cl-_0.1M_3c_60s and CdS-

Ti/TiO2_SO42-_0.1M_3c_60s. As shown in Fig. 6b, the concentration of the Cd precursor had no

f
significant effect on the photocurrent generated by the photoelectrodes. Despite the increase in

oo
the Cd precursor concentration, which in turn increased the amount of deposited CdS, the

pr
photocurrent did not increase. In terms of the number of SILAR cycles (Fig. 6c), it was observed

that the smallest photocurrent at about 0.5 mA cm-2 was generated by CdS-Ti/TiO2_NO3-
e-
_0.1M_5c_60s which underwent 5 cycles, and the best result was obtained for CdS-
Pr
Ti/TiO2_NO3-_0.1M_3c_60s, which underwent 3 cycles. As presented in Fig. 6d, longer lengths

of dipping time for the preparation of the photoelectrodes caused them to generate lower
al

photocurrents. Our results were consistent with data presented by Li et al. [28], where the
n

presence of CdS nanoparticles on the TiO2 surface caused the generation of a higher
ur

photocurrent than in the case of pristine TiO2. Moreover, the authors observed that the longer
Jo

electrodeposition time and the greater amount of CdS on TiO2, the generated photocurrent

increases, but only up to a certain point. Too long deposition time and thus the amount of CdS

may limit the generation of photocurrents due to the formation of charge recombination sites.

Additionally, for all the prepared photoelectrodes, linear voltammograms were obtained in the

dark (Fig. A2 (Supplementary data)). The results indicate that the CdS sensitization of Ti/TiO2

had a minor effect on the overpotential of oxygen generation from water at the photoelectrodes,

which was about 1.5 V.

15
3.1.6. Hydroxyl radical measurements

The amount of hydroxyl radicals generated during PEC was determined using the fluorescence

measurement of 2-hydroxyterephtalic acid, which forms from the reaction of hydroxyl radicals

with terephtalic acid. The fluorescence intensity of the 2-hydroxyterephtalic acid formed during

the PEC using Ti/TiO2 and CdS-Ti/TiO2 photoanodes was proportional to the amount of

hydroxyl radicals, as presented in Fig. A3 (Supplementary data). In the PEC process, more

f
oo
hydroxyl radicals were produced when CdS-Ti/TiO2 was used as the photoanode rather than with

Ti/TiO2. This indicates that CdS effectively increases the generation of hydroxyl radicals during

pr
PEC under UV-Vis light irradiation. This is most likely a result of the limitation of the

recombination of e-/h+ pairs and the generation of more electrons, which can contribute to the
e-
generation of hydroxyl radicals by transforming the superoxide anions formed by reducing
Pr
oxygen in the reaction with e-. The CdS-Ti/TiO2_Cl-_0.1M_3c_60s photoanode deviated from

the trend, as it yielded a similar amount of hydroxyl radicals to the Ti/TiO2 photoanode. For the
al

CdS-Ti/TiO2_NO3-_0.1M_3c_60s, CdS-Ti/TiO2_NO3-_0.1M_2c_60s, CdS-Ti/TiO2_NO3-


n

_0.1M_5c_60s, CdS-Ti/TiO2_NO3-_0.05M_3c_60s, CdS-Ti/TiO2_NO3-_0.1M_3c_30s, and CdS-


ur

Ti/TiO2_NO3-_0.1M_3c_120s photoanodes, comparable amounts of hydroxyl radicals after 3


Jo

hours of PEC were obtained, and these were much higher than those obtained in the case of

Ti/TiO2. Therefore, using CdS-Ti/TiO2_NO3-_0.2M_3c_60s and CdS-Ti/TiO2_ SO42-

_0.1M_3c_60s in PEC generates more hydroxyl radicals compared to Ti/TiO2, but less when

compared to the other CdS-Ti/TiO2 photoanodes. However, the results show no correlation

between the effectiveness of the generated hydroxyl radicals and PEC activity, which indicates

that not only radicals but also other oxidants are responsible for the degradation of pollutants in

16
PEC with CdS-Ti/TiO2. Participation in the PEC IF degradation of various oxidants is discussed

in the further section.

3.2. Photoelectrochemical activity of CdS-Ti/TiO2

The efficiency of PEC degradation strongly depends on the material and the properties of the

photoanode, which are determined by the parameters of its synthesis. Nanocomposite CdS-

Ti/TiO2 photoanodes were prepared using the SILAR method with varying parameters. CdS-

f
oo
Ti/TiO2_Cl-_0.1M_3c_60s, CdS-Ti/TiO2_SO42-_0.1M_3c_60s, and CdS-Ti/TiO2_NO3-

_0.1M_3c_60s were prepared with different Cd precursors. CdS-Ti/TiO2_NO3-_0.1M_3c_60s,

pr
CdS-Ti/TiO2_NO3-_0.05M_3c_60s, and CdS-Ti/TiO2_NO3-_0.2M_3c_60s were prepared with
e-
different Cd precursor concentrations. CdS-Ti/TiO2_NO3-_0.1M_3c_60s, CdS-Ti/TiO2_NO3-

_0.1M_2c_60s, and CdS-Ti/TiO2_NO3-_0.1M_5c_60s were prepared with different numbers of


Pr
SILAR cycles. Finally, CdS-Ti/TiO2_NO3-_0.1M_3c_60s, CdS-Ti/TiO2_NO3-_0.1M_3c_30s,

and CdS-Ti/TiO2_NO3-_0.1M_3c_120s were prepared with different times of dipping. The PEC
al

(Table 2) and PC (Table A3 (Supplementary data)) efficiency of the prepared photoanodes were
n

investigated in terms of their IF degradation rate and TOC removal. In the PEC and PC
ur

processes, a higher IF degradation efficiency was obtained for the CdS-Ti/TiO2 photoanode than
Jo

Ti/TiO2. Additionally, for each photoanode, the MCE of PEC was calculated (Table 2).

3.2.1. Effect of Cd precursor type and concentration

To synthesize CdS-Ti/TiO2_Cl-_0.1M_3c_60s, CdS-Ti/TiO2_SO42-_0.1M_3c_60s, and CdS-

Ti/TiO2_NO3-_0.1M_3c_60s, Cd precursor solutions of CdCl2 (0.1 M), CdSO4 (0.1 M), and

Cd(NO3)2 (0.1 M) were used, respectively. The most frequently chosen Cd precursor for the CdS

synthesis by the SILAR method was Cd(NO3)2 [9,14,26]. To the best of our knowledge, the

17
influence of the type of Cd precursor on the PEC activity of CdS-TiO2 nanocomposites has not

been studied so far. The influence of the type of Cd precursor on the PEC activity of the CdS-

Ti/TiO2 photoanodes is presented on Fig. 7a. The highest degradation rate constant was obtained

for CdS-Ti/TiO2_SO42-_0.1M_3c_60s (k = 0.0185 min-1), followed by CdS-Ti/TiO2_NO3-

_0.1M_3c_60s (k = 0.0184 min-1) and CdS-Ti/TiO2_Cl-_0.1M_3c_60s (k = 0.0159 min-1). TOC

removal was the most efficient when CdS-Ti/TiO2_NO3-_0.1M_3c_60s (48.1%) was employed.

f
For CdS-Ti/TiO2_Cl-_0.1M_3c_60s and CdS-Ti/TiO2_SO42-_0.1M_3c_60s, ΔTOC was 28.9%

oo
and 44.6% respectively. The increase in the efficiency of TOC removal increased the MCE

pr
(Table 2). The relatively low degradation efficiency of CdS-Ti/TiO2_Cl-_0.1M_3c_60s may be

due to the large amount and large size, over 5 nm, of the CdS QDs deposited on its surface, as
e-
confirmed by the XPS analysis (Table A2 (Supplementary data)). As seen on SEM image (Fig.
Pr
1), when CdCl2 was used as the Cd precursor, the CdS nanoparticles deposited on TiO2 formed

agglomerates, which may inhibit electron transport in the nanocomposite, as previously


al

described by Mustakin et al. [13]. In addition, in the same electrode, a slight absorption in the
n

visible light was observed, and it was found that it generates a minor amount of hydroxyl
ur

radicals which translates into its low IF decomposition efficiency. The photoelectrodes
Jo

synthesized using Cd(NO3)2 and CdSO4 showed a similar PEC activity and were characterized

by a similar amount and size of CdS QDs at 3.1 and 3.6 nm, respectively (Table A2

(Supplementary data)).

To investigate the effect of the Cd precursor concentration on the PEC degradation efficiency of

CdS-Ti/TiO2, Cd(NO3)2 solutions with concentrations of 0.05 M, 0.1 M, and 0.2 M were used to

fabricate the photoelectrodes. As expected, the increase in the concentration of the Cd precursor

increased the amount and size of the CdS QDs on the Ti/TiO2 (Table A2 (Supplementary data)),

18
it is in line with data presented by Mustakin et al. [13]. However, there was no correlation

between the amount of deposited CdS and the PEC activity of the nanocomposites. As presented

in Fig. 7b, CdS-Ti/TiO2_NO3-_0.2M_3c_60s was prepared with the highest concentration of Cd

precursor (0.2 M), but it exhibited a relatively low PEC efficiency. This may be the result of its

low visible irradiation absorption, which causes the generation of less photocurrent and hydroxyl

radicals. The PEC degradation of CdS-Ti/TiO2_NO3-_0.2M_3c_60s (k = 0.0075 min-1) was

f
slower than that of CdS-Ti/TiO2_NO3-_0.05M_3c_60s (k = 0.0091 min-1), which was prepared

oo
with the least concentration of Cd precursor (0.05 M). Nevertheless, compared with CdS-

pr
Ti/TiO2_NO3-_0.05M_3c_60s, the PEC degradation and TOC removal of CdS-Ti/TiO2_NO3-

_0.1M_3c_60s (k = 0.0184 min-1), prepared with 0.1 M Cd(NO3)2, was more than twice as fast
e-
(Table 2). For this electrode, the highest absorbance shift towards the visible light range was
Pr
observed. Similar observations were presented by Jun et al. [29]. There were prepared

nanocomposites CdS-TiO2 by SILAR method using Cd concentration in range 0.01 M – 1.0 M,


al

and applied as electrodes in quantum dot-sensitized solar cells. The highest solar cell efficiency
n

was obtained for photoelectrode prepared with 0.1 M precursors concentration. High precursors
ur

concentration causes more QDs generated, which results in their aggregation, which in turn
Jo

interferes with the transport of electrons.

3.2.2. Effect of number of SILAR cycles and dipping time

The number of SILAR cycles and dipping time are important parameters which can determine

the amount and size of the deposited QDs and the PEC activity. As expected, the more SILAR

cycles carried out, the larger the quantity and size of the CdS QDs on the Ti/TiO2 photoelectrode

(Table A2 (Supplementary data)). However, as the number of cycles increased, visible light

absorbance decreased. The CdS-Ti/TiO2_NO3-_0.1M_3c_60s electrode which underwent 3

19
SILAR cycles had the highest IF degradation rate and TOC removal (Table 2). A marginally

slower degradation was observed for the CdS-Ti/TiO2_NO3-_0.1M_2c_60s electrode which

underwent 2 cycles (Fig. 7c). However, increasing the number of cycles to 5 resulted in the

significant decrease in the photoelectrochemical activity of the electrode. Zhu et al. [30] reported

that too many cycles of the SILAR method limit the PC activity of CdS-Ti/TiO2 nanocomposites

due to the aggregation of the CdS QD which block the pores of the TiO2 NTs. This reduces the

f
access of reagents and light to the surface of the TiO2 NTs. Qorbani et al. [31] also indicated the

oo
greater possibility of recombination or blocking of electrons and holes when there are large

pr
amounts of CdS nanoparticles in the nanocomposite. It is worth noting that for the CdS-Ti/TiO2

photoelectrodes, PEC degradation efficiency is correlated to the intensity of the photocurrent


e-
generated by the photoelectrodes.
Pr
The comparison of the PEC efficiency of the CdS-Ti/TiO2 photoanodes prepared with different

times of dipping in the precursor solution are presented in Fig 7d. As the dipping time increased,
al

the amount and size of the CdS QDs deposited on the Ti/TiO2 electrode and its absorbance in the
n

visible light range increased, while photocurrent generation decreased. However, no correlation
ur

with the PEC activity was observed. The photoanodes prepared at the shortest (30 s) and the
Jo

longest times (120 s) of dipping showed a lower efficiency in IF degradation and TOC removal

than the CdS-Ti/TiO2_NO3-_0.1M_3c_60s electrode, with a dipping time of 60 s. Similar to the

results of the observation on the number of cycles, a dipping time which is too long results in the

deposition of a large amount of CdS, which limits the PEC activity.

In summary, the highest efficiency of PEC IF degradation and TOC removal were obtained for

the CdS-Ti/TiO2_NO3-_0.1M_3c_60s photoelectrode, which was prepared using 0.1 M

Cd(NO3)2 solution as the Cd precursor and 3 SILAR cycles with 60 s of dipping time. The

20
SILAR parameter that had the greatest impact on PEC activity of the CdS-Ti/TiO2

photoelectrodes was the concentration of the Cd precursor. Nevertheless, the number of cycles

and the dipping time also had a significant effect on PEC activity. It was also observed that as

the number of deposited CdS QDs increases, their size increases, and QDs with sizes below 4 nm

more effectively enhance PEC activity than larger CdS nanoparticles.

f
oo
3.3. Comparison of photocatalytic, electrochemical, and photoelectrocatalytic activity of

CdS-Ti/TiO2

pr
The comparison of electrolysis (applied potential of 1 V), photocatalysis (UV-Vis irradiation at
e-
550 W m-2), and photoelectrocatalysis (applied potential of 1 V, UV-Vis irradiation at 550 W m-
Pr
2
) of IF using Ti/TiO2 and the CdS-Ti/TiO2_NO3-_0.1M_3c_60s photoanode were carried out. In

addition, the photolysis of IF under UV-Vis irradiation (550 W m-2) was investigated. As shown
al

in Fig. 8, there was no significant IF degradation under the photolysis and electrolysis with both
n

photoanodes. IF was not directly photolyzed because it does not absorb UV-Vis irradiation due
ur

to the lack of aromatic rings or double C-C bonds in its structure, which was also verified by

other authors [32–34]. On the other hand, the electrochemical degradation of IF did not take
Jo

place due to the system's operation at low potential (1 V) at which no generation of hydroxyl

radicals was observed. These observations were consistent with the results presented earlier [19].

PC degradation was a pseudo first order reaction with a degradation rate of 0.0044 min-1 (R2 =

0.9857) for Ti/TiO2 and 0.0083 min-1 (R2 = 0.9975) for Ti/TiO2_NO3-_0.1M_3c_60s, which

indicates that the presence of CdS QDs increases the PC activity of the photoanodes. An increase

PC activity of TiO2 due to the presence of CdS nanoparticle was also observed by Yu et al. [16]

and Kalarivalappil et al. [9] for degradation of rhodamine B and methyl orange. The higher PC

21
activity of the Ti/TiO2_NO3-_0.1M_3c_60s photoelectrode explains its higher TOC removal at

8%, which was not observed in Ti/TiO2. The same dependence was observed in the

photoelectrocatalysis, where the degradation and the TOC removal of CdS-Ti/TiO2_NO3-

_0.1M_3c_60s (k = 0.0184 min-1, R2 = 0.9916; ΔTOC = 48%) was 2.5 times faster and higher by

11% than that of Ti/TiO2 (k = 0.0072 min-1, R2 = 0.9827; ΔTOC = 37%), respectively. Based on

the above-mentioned results, the higher IF removal efficiency in photoelectrocatalysis than

f
photocatalysis was observed for both the Ti/TiO2 and CdS-Ti/TiO2_NO3-_0.1M_3c_60s

oo
photoanodes. The same was observed for the other CdS-Ti/TiO2 photoanodes (Table A3

pr
(Supplementary data)). Obtained results using other photoelectrodes, in agreement with our

studies were presented by more authors [35–37]. This is because the applied potential during
e-
photocatalysis is not sufficient to induce the recombination of the charge carriers, while the
Pr
recombination of carriers takes place during photocatalysis, inhibiting the generation of oxidants.

Moreover, the increase in the efficiency in photoelectrocatalysis relative to photocatalysis can be


al

a result of the additional generation of oxidants on the cathode [38]. A precise description of the
n

PEC mechanism is presented in following section. The electric energy required (EEO) for the
ur

PEC and PC degradation of IF for both the Ti/TiO2 and CdS-Ti/TiO2_NO3-_0.1M_3c_60s

photoelectrodes were calculated. For CdS-Ti/TiO2_NO3-_0.1M_3c_60s, the EEO values for the
Jo

PEC and PC degradation were 5.37 and 12.5 kWh m-3 order-1, respectively. For Ti/TiO2 the

electric energy required for PEC degradation (EEO = 13.76 kWh m-3 order-1) was lower than that

for PC degradation (EEO = 25.7 kWh m-3 order-1). These results show that, regardless of the

electrode employed, a smaller amount of electric energy is required for the removal IF in the

PEC process than in the PC process. Additionally, the amount of electric energy used for the

PEC process with CdS-Ti/TiO2_NO3-_0.1M_3c_60s was more than twice lower than that with

22
Ti/TiO2. Therefore, it can be concluded that photoelectrocatalysis with CdS-Ti/TiO2_NO3-

_0.1M_3c_60s is more cost effective and suitable for practical applications.

3.4. Mechanism of photoelectrocatalytic degradation

PEC degradation is based on the generation of active species (such as ∙OH, O2-∙ and h+) from

Reactions (2)–(10). The type of active species involved in the PEC degradation of organic

compounds depend on the photoanode material and the type of pollutant [19]. For the TiO2

f
oo
photoanode under UV irradiation, an e-/h+ pair is formed (Reaction (2)). By applying an external

potential, e- is transported from the conduction band of TiO2 to the cathode via an external

pr
circuit. In this manner, the recombination of the photogenerated e-/h+ pairs is limited. The h+
e-
staying in the TiO2 valence band can convert H2O molecules or OH- to highly reactive hydroxyl
Pr
radicals (Reactions (4) and (5)), which are strong oxidants that can degrade a wide range of

organic compounds. Some of the ∙OH diffuse into the liquid phase, and the remaining stays

adsorbed onto the photoanode surface. Moreover, a h+ can oxidize the organic compounds
al

(Reaction (6)) adsorbed onto the photoanode surface. The e- transported to the cathode can
n

reduce oxygen molecules to O2-∙ (Reaction (7)) and HO2∙ (Reaction (8)). Further, H2O2 can be
ur

formed according to Reaction (9), and H2O2 could be converted into ∙OH, according to Reactions
Jo

(10) and (11). For the TiO2 sensitized with CdS QDs, except for the reactions on TiO2, under

UV-Vis irradiation, e- from the CdS valence band transfers to the CdS conduction band

(Reaction (3)) and then to the TiO2 conduction band, reducing the recombination of e-/h+ pairs.

This is possible because of the more negative position of the conduction band of CdS than that of

TiO2 [9]. The valence band of CdS is at a lower value than that of TiO2, therefore, h+ could be

moved to the valence band of CdS. The general mechanism for the PEC degradation of organic

pollutants using a CdS-Ti/TiO2 photoelectrode is presented on Fig. 9.

23
𝑇𝑖𝑂2 + ℎ𝑣 → ℎ+ + 𝑒 − (2)

𝐶𝑑𝑆 − 𝑇𝑖𝑂2 + ℎ𝑣 → ℎ+ + 𝑒 − (3)

ℎ+ + 𝐻2 𝑂 → 𝑂𝐻 ∙ +𝐻 + (4)

ℎ+ + 𝑂𝐻− → 𝑂𝐻 ∙ (5)

ℎ+ + 𝑅 → 𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒𝑠 → 𝐶𝑂2 + 𝐻2 𝑂 (6)

f
oo
𝑒 − + 𝑂2 → 𝑂2− ∙ (7)

𝑒 − + 𝑂2 + 𝐻 + → 𝐻𝑂𝑂 ∙

pr
(8)

𝑒 − + 𝑂2−∙ + 2 𝐻 + → 𝐻2 𝑂2 (9)
e-
𝐻2 𝑂2 + 𝑂2− ∙ → 𝑂𝐻 ∙ +𝑂𝐻− + 𝑂2 (10)
Pr
𝐻2 𝑂2 + ℎ𝑣 → 2 𝑂𝐻 ∙ (11)
al

To compare the contribution of the active species in the PEC degradation of IF with Ti/TiO2 and
n

CdS-Ti/TiO2_NO3-_0.1M_3c_60s, the scavengers of the main oxidants was used. The results of
ur

the PEC degradation with the scavengers are presented in Fig. 10. Methanol (MeOH) was added
Jo

to the treated IF solution to inhibit the activity of the ∙OH generated in Reactions (4), (5), and

reaction (10). The h+ in the valence band of TiO2 and CdS and the h+ generated from ∙OH were

blocked using formic acid (FA). Ascorbic acid (AA) inhibited the action of O2-∙ and the

additional ∙OH formed in Reactions (9) and (10). The simultaneous usage of MeOH and FA

inhibits the activity of both the h+ and the free ∙OH.

The results imply that the ∙OH were the dominant species responsible for the IF degradation in

both the Ti/TiO2 and CdS-Ti/TiO2_NO3-_0.1M_3c_60s photoanodes. For Ti/TiO2, the IF

24
degradation of 70% is the result of the oxidation by the free ∙OH generated during the reaction

with H2O or OH- with h+ (around 60%) and the transformation of O2-∙ (Reactions (9) and (10))

(around 10%). For CdS-Ti/TiO2_NO3-_0.1M_3c_60s, the contribution of ∙OH was 55%, more of

which generated is by h+ (around 44%). The other species responsible for IF degradation was h+;

its precipitation during the IF oxidation was higher when CdS-Ti/TiO2_NO3-_0.1M_3c_60s

(around 24%) was employed rather than Ti/TiO2 (around 13%). This is probably is the result of

f
the higher amount of h+ forming not only on TiO2 but also on CdS (Reaction (4)). Moreover, the

oo
h+ in the CdS valence band cannot oxidize H2O and OH- to ∙OH because of the position of the

pr
valence band. The other oxidative species generated during the photoelectrocatalysis contributed

to 22% and 13% of the IF degradation when using CdS-Ti/TiO2_NO3-_0.1M_3c_60s and


e-
Ti/TiO2, respectively, and it has a similar participation as h+. These results indicate that CdS QDs
Pr
on TiO2 improves the PEC IF degradation efficiency through the generation of h+ and other

oxidative species.
al

3.5. Photoelectrocatalytic IF degradation pathway


n

The PEC degradation pathway of IF with both the Ti/TiO2 and CdS-Ti/TiO2_NO3-_0.1M_3c_60s
ur

photoanodes were proposed based on the transformation products identified using liquid
Jo

chromatography mass spectrometry. The samples for analysis were collected after 1, 2, and 3 h

of PEC degradation. As shown in Fig. 11, the products with [M+H]+ = 277; 275; 259; 279; and

265 where generated in the photoelectrocatalysis with both photoanodes. Additionally, in the

case of Ti/TiO2, compounds with [M+H]+ = 199 and 253 were detected. The [M+H]+= 277

product, which was identified as 4-hydroxy-ifosfamide, likely formed from the ∙OH attack at the

C4 position. 4-hydroxy-ifosfamide can be converted into 4-ketoifosfamide ([M+H]+ = 275)

through keto-enol tautomerization or into imino-ifosfamide ([M+H]+= 259) through

25
dehydroksylation. Previously, other authors have reported the formation of 4-hydroxy-

ifosfamide, 4-ketoifosfamide, and iminoifosfamide as degradation products of the

electrochemical, photochemical, and PC oxidation of IF [21,39,40]. In the next step of the IF

PEC decomposition, the opening of the cyclic ring was observed, as evidenced by products

[M+H]+ = 279 and [M+H]+ = 265. During the PEC degradation with Ti/TiO2, additional products

with [M+H]+ = 199 and [M+H]+ = 253 were detected. The product with [M+H]+ = 199 is N-

f
deschloroethyl-ifosfamide, which arose from the detachment of one of the aliphatic chains. It has

oo
already been shown as an intermediate formed through the photochemical oxidation of IF [40].

pr
The product with [M+H]+ = 253 was created via chlorine substitution. The further decomposition

of IF leads to the release of inorganic ions and CO2 from low molecular weight organic
e-
compounds, such as carboxylic acids.
Pr
3.6. Conclusions
al

As presented research has shown, the parameters of the SILAR method affect both the

efficiency of the PEC degradation and the optical and photoelectrochemical properties of the
n

nanocomposite, which is related to the amount and size of the CdS QDs. Using different
ur

types of Cd precursors, it was found that using CdCl2 increased the amount of CdS deposited
Jo

by almost ten times compared to using CdSO4 and Cd(NO3)2. A larger amount of CdS was

associated with the formation of larger sizes of QDs and their aggregation on the surface of

the TiO2 NTs which resulted in the lowest efficiency of IF degradation and TOC removal.

The increase in the concentration of the Cd precursor, the number of SILAR cycles and

dipping time increased the amount and size of the CdS QDs deposited on Ti/TiO2. The

highest efficiency of PEC degradation of IF and TOC removal was exhibited by the CdS-

Ti/TiO2 photoelectrode, which was prepared using 0.1 M Cd(NO3)2 as the Cd precursor and

26
at 3 SILAR cycles with a dipping time of 60 s. Comparing the PEC degradation of pristine

Ti/TiO2 and CdS-Ti/TiO2 showed that the nanocomposite had a higher PC and PEC activity

due to the absence of photolysis and electrochemical oxidation of IF. This indicates that the

presence of CdS QDs increased the PC activity of CdS-Ti/TiO2 by increasing activity under

the visible radiation. In addition, studies on the mechanism of the PEC degradation of IF in

the presence of CdS-Ti/TiO2 have shown that h+ and other oxidants had a greater

f
contribution to the degradation than when pristine Ti/TiO2 was used. Finally, the PEC IF

oo
degradation pathway based on identified organic degradation products was proposed. The

pr
decomposition of IF occurred initially via cyclic ring hydroxylation and then opening.

Ultimately, mineralization and release of inorganic ions occurred.


e-
Funding:
Pr
This work was supported by the Polish National Science Center under the “Nanocomposites

Ti/TiO2(NTs)/XaSb as electrode materials in photoelectrochemical oxidation processes” grant


al

[grant number 2015/19/D/ST5/00710].


n
ur

References
Jo

[1] Q. Zhou, Z. Fang, J. Li, M. Wang, Applications of TiO 2 nanotube arrays in

environmental and energy fields : A review, Microporous Mesoporous Mater. 202 (2015)

22–35. https://doi.org/10.1016/j.micromeso.2014.09.040.

[2] P. Mazierski, M. Nischk, M. Go??kowska, W. Lisowski, M. Gazda, M.J. Winiarski, T.

Klimczuk, A. Zaleska-Medynska, Photocatalytic activity of nitrogen doped TiO2

nanotubes prepared by anodic oxidation: The effect of applied voltage, anodization time

and amount of nitrogen dopant, Appl. Catal. B Environ. 196 (2016) 77–88.

27
https://doi.org/10.1016/j.apcatb.2016.05.006.

[3] L. Han, Y. Xin, H. Liu, X. Ma, G. Tang, Photoelectrocatalytic properties of nitrogen

doped TiO2/Ti photoelectrode prepared by plasma based ion implantation under visible

light, J. Hazard. Mater. 175 (2010) 524–531.

https://doi.org/10.1016/j.jhazmat.2009.10.037.

[4] Y. Su, S. Han, X. Zhang, X. Chen, L. Lei, Preparation and visible-light-driven

f
oo
photoelectrocatalytic properties of boron-doped TiO2 nanotubes, Mater. Chem. Phys. 110

(2008) 239–246. https://doi.org/10.1016/j.matchemphys.2008.01.036.

pr
[5] A.M. Husin, L. Jeffery, M.B. Kassim, Carbon doped TiO 2 nanotubes photoanodes
e-
prepared by in-situ anodic oxidation of Ti-foil in acidic and organic medium with
Pr
photocurrent enhancement, Ceram. Int. 39 (2013) 3731–3739.

https://doi.org/10.1016/j.ceramint.2012.10.209.
al

[6] X. Cheng, H. Liu, Q. Chen, J. Li, P. Wang, Preparation and characterization of palladium
n

nano-crystallite decorated TiO 2 nano-tubes photoelectrode and its enhanced


ur

photocatalytic efficiency for degradation of diclofenac, J. Hazard. Mater. 254–255 (2013)


Jo

141–148. https://doi.org/10.1016/j.jhazmat.2013.03.062.

[7] H. Wang, W. Liang, W. Zhang, D. Zhou, Preparation and photoelectric properties of Pt /

TiO 2 nanotube electrodes by a pre-doping method, Thin Solid Films. 653 (2018) 101–

106. https://doi.org/10.1016/j.tsf.2018.03.030.

[8] M. Qorbani, N. Naseri, O. Moradlou, R. Azimirad, A.Z. Moshfegh, Applied Catalysis B :

Environmental How CdS nanoparticles can influence TiO 2 nanotube arrays in solar

energy applications ?, 162 (2015) 210–216.

28
[9] V. Kalarivalappil, S.J. Hinder, S.C. Pillai, V. Kumar, B.K. Vijayan, Stability studies of

CdS sensitized TiO2 nanotubes prepared using the SILAR method, J. Environ. Chem.

Eng. 6 (2018) 1404–1413. https://doi.org/10.1016/j.jece.2018.01.050.

[10] D. Esparza, I. Zarazúa, T. López-Luke, R. Carriles, A. Torres-Castro, E.D. La Rosa,

Photovoltaic Properties of Bi2S3 and CdS Quantum Dot Sensitized TiO2 Solar Cells,

Electrochim. Acta. 180 (2015) 486–492. https://doi.org/10.1016/j.electacta.2015.08.102.

f
oo
[11] Z. Liu, B. Wang, J. Wu, Q. Dong, X. Zhang, H. Xu, Effects of Hydroxylation on PbS

Quantum Dot Sensitized TiO2 Nanotube Array Photoelectrodes, Electrochim. Acta. 187

pr
(2016) 480–487. https://doi.org/10.1016/j.electacta.2015.11.042.
e-
[12] D. Zhao, C. Yang, Recent advances in the TiO 2 / CdS nanocomposite used for
Pr
photocatalytic hydrogen production and quantum-dot-sensitized solar cells, Renew.

Sustain. Energy Rev. 54 (2016) 1048–1059. https://doi.org/10.1016/j.rser.2015.10.100.


al

[13] N.S.M. Mustakim, C.A. Ubani, S. Sepeai, N.A. Ludin, M.A.M. Teridi, M.A. Ibrahim,
n

Quantum dots processed by SILAR for solar cell applications, Sol. Energy. 163 (2018)
ur

256–270. https://doi.org/10.1016/j.solener.2018.02.003.
Jo

[14] Z. Lan, W. Wu, S. Zhang, L. Que, J. Wu, Preparation of high-ef fi ciency CdS quantum-

dot-sensitized solar cells based on ordered TiO 2 nanotube arrays, Ceram. Int. 42 (2016)

8058–8065. https://doi.org/10.1016/j.ceramint.2016.02.003.

[15] Y. Zhu, Y. Wang, Z. Chen, L. Qin, L. Yang, L. Zhu, P. Tang, T. Gao, Y. Huang, Z. Sha,

G. Tang, Visible light induced photocatalysis on CdS quantum dots decorated TiO2

nanotube arrays, Appl. Catal. A Gen. 498 (2015) 159–166.

https://doi.org/http://dx.doi.org/10.1016/j.apcata.2015.03.035.

29
[16] L. Yu, D. Wang, D. Ye, CdS nanoparticles decorated anatase TiO 2 nanotubes with

enhanced visible light photocatalytic activity, Sep. Purif. Technol. 156 (2015) 708–714.

https://doi.org/10.1016/j.seppur.2015.10.069.

[17] J. Zhang, V.W.C. Chang, A. Giannis, J. Wang, Removal of cytostatic drugs from aquatic

environment : A review, Sci. Total Environ. 445–446 (2013) 281–298.

https://doi.org/10.1016/j.scitotenv.2012.12.061.

f
oo
[18] A. Białk-Bielińska, E. Mulkiewicz, M. Stokowski, S. Stolte, P. Stepnowski, Acute aquatic

toxicity assessment of six anti-cancer drugs and one metabolite using biotest battery –

pr
Biological effects and stability under test conditions, Chemosphere. 189 (2017) 689–698.
e-
https://doi.org/10.1016/j.chemosphere.2017.08.174.
Pr
[19] P. Mazierski, A. Fiszka Borzyszkowska, P. Wilczewska, A. Białk-Bielinska, A. Zaleska-

Medynska, E.M. Siedlecka, A. Pieczy?nska, Removal of 5- fl uorouracil by solar-driven


al

photoelectrocatalytic oxidation using Ti / TiO 2 ( NT ) photoelectrodes, Water Res. 157


n

(2019) 610–620. https://doi.org/10.1016/j.watres.2019.04.010.


ur

[20] P. Mazierski, J. Nadolna, W. Lisowski, M.J. Winiarski, M. Gazda, M. Nischk, T.


Jo

Klimczuk, A. Zaleska-Medynska, Effect of irradiation intensity and initial pollutant

concentration on gas phase photocatalytic activity of TiO2 nanotube arrays, Catal. Today.

284 (2017) 19–26. https://doi.org/10.1016/j.cattod.2016.09.004.

[21] M.E. Siedlecka, A. Ofiarska, A. Fiszka Borzyszkowska, A. Białk-Bielińska, P.

Stepnowski, A. Pieczyńska, Cytostatic drug removal using electrochemical oxidation with

BDD electrode : Degradation pathway and toxicity, Water Res. 144 (2018) 235–245.

https://doi.org/10.1016/j.watres.2018.07.035.

30
[22] S. Garcia-Segura, H.O.N. Tugaoen, K. Hristovski, P. Westerho, Photon flux influence on

photoelectrochemical water treatment, Electrochem. Commun. 87 (2018) 63–65.

https://doi.org/10.1016/j.elecom.2017.12.026.

[23] A. V. Naumkin, A. Kraut-Vass, S.W. Gaarenstroom, C.J. Powell, NIST X-ray

Photoelectron Spectroscopy Database, NIST Stand. Ref. Database. 20 (2012).

[24] J. Nanda, B.A. Kuruvilla, D.D. Sarma, Photoelectron spectroscopic study of CdS

f
oo
nanocrystallites, Phys. Rev. B. 59 (1999) 7473–7479.

pr
[25] U. Winkler, D. Eich, Z.H. Chen, R. Fink, S.K. Kulkarni, E. Umbach, Detailed

investigation of CdS nanoparticle surfaces by high-resolution photoelectron spectroscopy,


e-
Chem. Phys. Lett. 306 (1999) 95–102. https://doi.org/10.1016/S0009-2614(99)00427-3.
Pr
[26] S.A. Pawar, D.S. Patil, A.C. Lokhande, M.G. Gang, J.C. Shin, P.S. Patil, J.H. Kim,

Chemical synthesis of CdS onto TiO2 nanorods for quantum dot sensitized solar cells,
al

Opt. Mater. (Amst). 58 (2016) 46–50. https://doi.org/10.1016/j.optmat.2016.05.019.


n

[27] X. Deng, H. Zhang, R. Guo, Y. Cui, Q. Ma, X. Zhang, X. Cheng, B. Li, M. Xie, Q.
ur

Cheng, Effect of fabricating parameters on photoelectrocatalytic performance of


Jo

CeO2/TiO2 nanotube arrays photoelectrode, Sep. Purif. Technol. 193 (2018) 264–273.

https://doi.org/10.1016/j.seppur.2017.10.066.

[28] G. Li, L. Wu, F. Li, P. Xu, D. Zhang, H. Li, Photoelectrocatalytic degradation of organic

pollutants via a CdS quantum dots enhanced TiO2 nanotube array electrode under visible

light irradiation, Nanoscale. 5 (2013) 2118–2125. https://doi.org/10.1039/c3nr34253k.

[29] H.K. Jun, M.A. Careem, A.K. Arof, Fabrication, characterization, and optimization of

31
CdS and CdSe quantum Dot-sensitized solar cells with quantum dots prepared by

successive ionic layer adsorption and reaction, Int. J. Photoenergy. 2014 (2014).

https://doi.org/10.1155/2014/939423.

[30] Y. Zhu, Y. Wang, Z. Chen, L. Qin, L. Yang, L. Zhu, Applied Catalysis A : General

Visible light induced photocatalysis on CdS quantum dots decorated TiO 2 nanotube

arrays, "Applied Catal. A, Gen. 498 (2015) 159–166.

f
oo
https://doi.org/10.1016/j.apcata.2015.03.035.

[31] M. Qorbani, N. Naseri, O. Moradlou, R. Azimirad, A.Z. Moshfegh, How CdS

pr
nanoparticles can influence TiO2 nanotube arrays in solar energy applications?, Appl.
e-
Catal. B Environ. 162 (2015) 210–216. https://doi.org/10.1016/j.apcatb.2014.06.053.
Pr
[32] A. Ofiarska, A. Pieczyńska, A. Fiszka Borzyszkowska, P. Stepnowski, E.M. Siedlecka, Pt-

TiO<inf>2</inf>-assisted photocatalytic degradation of the cytostatic drugs ifosfamide


al

and cyclophosphamide under artificial sunlight, Chem. Eng. J. 285 (2016).


n

https://doi.org/10.1016/j.cej.2015.09.109.
ur

[33] M. Česen, T. Kosjek, M. Laimou-Geraniou, B. Kompare, B. Širok, D. Lambropolou, E.


Jo

Heath, Occurrence of cyclophosphamide and ifosfamide in aqueous environment and their

removal by biological and abiotic wastewater treatment processes, Sci. Total Environ. 528

(2015) 465–473. https://doi.org/10.1016/j.scitotenv.2015.04.109.

[34] A. Pieczyńska, A.F. Borzyszkowska, E.M. Siedlecka, Removal of cytostatic drugs by

AOPs : A review of applied processes in the context of green technology, Crit. Rev.

Environ. Sci. Technol. 47 (2017) 1282–1335.

https://doi.org/10.1080/10643389.2017.1370990.

32
[35] Z. Hua, Z. Dai, X. Bai, Z. Ye, P. Wang, H. Gu, X. Huang, Copper nanoparticles sensitized

TiO 2 nanotube arrays electrode with enhanced photoelectrocatalytic activity for

diclofenac degradation, Chem. Eng. J. 283 (2016) 514–523.

https://doi.org/10.1016/j.cej.2015.07.072.

[36] S. Chai, G. Zhao, P. Li, Y. Lei, Y.N. Zhang, D. Li, Novel sieve-like SnO2/TiO2

nanotubes with integrated photoelectrocatalysis: Fabrication and application for efficient

f
oo
toxicity elimination of nitrophenol wastewater, J. Phys. Chem. C. 115 (2011) 18261–

18269. https://doi.org/10.1021/jp205228h.

pr
[37] L. Zhang, Q. Shi, Y. Guo, D. Xu, H. Wang, L. Wang, Z. Bian, Interface optimization by
e-
impedance spectroscopy and photoelectrocatalytic degradation of clofibric acid,
Pr
Electrochim. Acta. 300 (2019) 242–252. https://doi.org/10.1016/j.electacta.2019.01.103.

[38] S. Garcia-Segura, E. Brillas, Applied photoelectrocatalysis on the degradation of organic


al

pollutants in wastewaters, J. Photochem. Photobiol. C Photochem. Rev. 31 (2017) 1–35.


n

https://doi.org/10.1016/j.jphotochemrev.2017.01.005.
ur

[39] W.W.-P. Lai, H.H.-H. Lin, A.Y.-C. Lin, TiO2 photocatalytic degradation and
Jo

transformation of oxazaphosphorine drugs in an aqueous environment., J. Hazard. Mater.

287C (2015) 133–141. https://doi.org/10.1016/j.jhazmat.2015.01.045.

[40] M. Česen, T. Kosjek, F. Busetti, B. Kompare, E. Heath, Human metabolites and

transformation products of cyclophosphamide and ifosfamide: analysis, occurrence and

formation during abiotic treatments, Environ. Sci. Pollut. Res. (2016).

https://doi.org/10.1007/s11356-016-6321-1.

33
Table 1. The preparation conditions of CdS-Ti/TiO2 photoelectrodes

SILAR parameters

Concentration
Photoelectrode Number of Time of SILAR
Cd precursor of Cd precursor
SILAR cycle cycle (s)
(M)

CdS-Ti/TiO2_Cl-_0.1M_3c_60s CdCl2 0.1 3 60

CdS-Ti/TiO2_SO42-_0.1M_3c_60s CdSO4

f
oo
CdS-Ti/TiO2_NO3-_0.1M_3c_60s Cd(NO3)2

CdS-Ti/TiO2_NO3-_0.05M_3c_60s 0.05

CdS-Ti/TiO2_NO3-_0.2M_3c_60s

pr
0.2

CdS-Ti/TiO2_NO3-_0.1M_2c_60s 0.1 2

CdS-Ti/TiO2_NO3-_0.1M_5c_60s
e- 5

CdS-Ti/TiO2_NO3-_0.1M_3c_30s 3 30
Pr
CdS-Ti/TiO2_NO3-_0.1M_3c_120s 120

Ti/TiO2 - - - -
n al
ur
Jo

34
Table 2. First rate constant k for IF PEC decomposition, TOC removal, and MCE.

Photoelectrode k ×103 (min-1) R2 ΔTOC (%) MCE (%)

CdS-Ti/TiO2_Cl-_0.1M_3c_60s 15.9 ± 0.4 0.9459 28.9 (±1.5) 73.4

CdS-Ti/TiO2_SO42-_0.1M_3c_60s 18.5 ± 0.1 0.9885 44.6 (±2.2) 113.2

CdS-Ti/TiO2_NO3-_0.1M_3c_60s 18.429 ± 0.001 0.9916 48.1 (±2.4) 122.1

CdS-Ti/TiO2_NO3-_0.05M_3c_60s 9.1 ± 0.4 0.9883 20.3 (±1.0) 89.7

CdS-Ti/TiO2_NO3-_0.2M_3c_60s 7.5 ± 0.4 0.9953 21.4 (±1.1) 49.2

f
CdS-Ti/TiO2_NO3-_0.1M_2c_60s 14.6 ± 0.5 0.9948 35.3 (±1.8) 98.1

oo
CdS-Ti/TiO2_NO3-_0.1M_5c_60s 8.9 ± 0.4 0.9545 38.6 (±1.9) 51.6

CdS-Ti/TiO2_NO3-_0.1M_3c_30s 11.9 ± 0.4 0.9760 34.2 (±1.4) 54.3

pr
CdS-Ti/TiO2_NO3-_0.1M_3c_120s 13.6 ± 0.1 0.9803 35.0 (±1.4) 87.0

Ti/TiO2 7.2 ± 0.2 e- 0.9827 37.2 (±1.5) 89.0


Pr
n al
ur
Jo

Fig. 1. Surface and cross-sectional morphology of prepared photoelectrodes.

35
f
oo
pr
e-
Pr
n al
ur
Jo

Fig. 2. TEM image of prepared photoelectrodes.

36
a) CdS-Ti/TiO2_Cl-_0.1M_3c_60s
b) CdS-Ti/TiO2_ NO3-_0.05M_3c_60s

CdS-Ti/TiO2_ SO42-_0.1M_3c_60s

CdS-Ti/TiO2_NO3-_0.1M_3c_60s

Intensity (a.u.)
Intensity (a.u.)

CdS-Ti/TiO2_NO3-_0.1M_3c_60s

CdS-Ti/TiO2_ NO3-_0.2M_3c_60s

Ti/TiO2

f
20 40 60 80 25 30 35

oo
20 40 60 80 25 30 35
2Q (deg.)
2Q (deg.)

pr
c) CdS-Ti/TiO2 _ NO3-_0.1M_2c_60s d) CdS-Ti/TiO2_ NO3-_0.1M_3c_30s

e-
Intensity (a.u.)

CdS-Ti/TiO2_NO3-_0.1M_3c_60s
CdS-Ti/TiO2_NO3-_0.1M_3c_60s
Intensity (a.u.)

Pr

CdS-Ti/TiO2_ NO3-_0.1M_5c_60s CdS-Ti/TiO2_ NO3-_0.1M_3c_120s


n al

20 40 60 80 25 30 35 20 40 60 80 25 30 35
2Q (deg.)
2Q (deg.)
ur
Jo

Fig. 3. pXRD patterns of the CdS-Ti/TiO2 photoanodes prepared using the SILAR method with various

parameters. a) Effect of precursor type and a reference sample (Ti/TiO2), b) effect of concentration of Cd

precursor, c) effect of number of SILAR cycles, and d) effect of dipping time. The red and black

vertical bars below the patterns represent the Bragg positions for TiO2 and Ti, respectively.

37
f
oo
pr
e-
Fig. 4. High-resolution S 2p XPS spectra of the CdS-Ti/TiO2 photoanodes prepared using the SILAR

method with various parameters. a) Effect of precursor type and a reference sample (Ti/TiO2), b) effect of
Pr
concentration of Cd precursor, c) effect of number of SILAR cycles, and d) effect of dipping time.
n al
ur
Jo

38
f
oo
pr
e-
Pr
al

Fig. 5. UV-Vis absorption spectra of CdS-Ti/TiO2 prepared using the SILAR method with various

parameters. a) Effect of precursor type, b) effect of concentration of Cd precursor, c) effect of number of


n

cycles, and d) effect of dipping time.


ur
Jo

39
f
oo
pr
e-
Pr
Fig. 6. Transient photocurrent response at 1 V of CdS-Ti/TiO2 prepared through the SILAR

method with various parameters. a) Effect of precursor type, b) effect of concentration of


al

Cd precursor, c) effect of number of cycles, and d) effect of dipping time.


n
ur
Jo

40
f
oo
pr
e-
Pr
n al
ur

Fig 7. Effect of a) type of Cd precursor: CdCl2, CdSO4, and Cd(NO3)2; b) concentration of Cd precursor:

0.05 M, 0.1 M, and 0.2 M; c) number of SILAR cycles: 2 cycles, 3 cycles, and 5 cycles; and d) dipping
Jo

time: 30 s, 60 s, 120 s on IF (20 mg L-1) PEC degradation efficiency with UV-Vis irradiation of 550 W m-
2
and potential of 1 V.

41
f
oo
Fig. 8. Photoelectrocatalytic (PEC), photocatalytic (PC), electrochemical (E), and photolytic (P)

pr
degradation of IF (20 mg L-1) using CdS-Ti/TiO2 and Ti/TiO2

e-
Pr
n al
ur
Jo

Fig. 9. General PEC degradation mechanism with CdS-Ti/TiO2 photoelectrode.

42
f
oo
Fig. 10. PEC degradation (550 W m-2, E = 1 V) of IF (20 mg L-1) on a) Ti/TiO2 and b) CdS-Ti/TiO2 with

scavengers: MeOH, FA, and AA.

pr
e-
Pr
n al
ur
Jo

Fig. 11. Proposed degradation pathway for IF during photoelectrocatalysis using Ti/TiO2 and CdS-

Ti/TiO2.

43
Graphical abstract

f
oo
pr
e-
Pr
n al
ur
Jo

44
CRediT author statement

Aleksandra Pieczyńska: Supervision, Investigation, Writing - Original Draft, Funding

acquisition, Project administration, Writing - Review & Editing

Paweł Mazierski: Investigation, Writing - Original Draft, Project administration

Wojciech Lisowski: Investigation,

f
oo
Tomasz Klimczuk: Investigation,

Adriana Zaleska-Medynska: Visualization, Writing - Review & Editing

pr
Ewa Siedlecka: Visualization, Writing - Review & Editing
e-
Pr
n al
ur
Jo

45
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

f
oo
pr
e-
Pr
n al
ur
Jo

46
Highlights
• The presence of CdS QDs in TiO2 enhances PEC degradation and TOC removal efficiency
• Precursor concentration, number of cycles, and time are the most key SILAR parameters
• The amount and size of CdS QDs determine the properties and PEC activity of electrode
• The presence of CdS increases the role of h+ and other oxidants in PEC mechanism

f
oo
pr
e-
Pr
n al
ur
Jo

47

You might also like