You are on page 1of 19

Research on Chemical Intermediates (2021) 47:5229–5247

https://doi.org/10.1007/s11164-021-04595-4

Facile synthesis of quantum dots/TiO2 photocatalyst


with superior photocatalytic activity: the effect of carbon
nitride quantum dots and N‑doped carbon dots

Ping Zhao1 · Bo Jin1 · Qingchun Zhang1 · Rufang Peng1

Received: 17 June 2021 / Accepted: 21 September 2021 / Published online: 18 October 2021
© The Author(s), under exclusive licence to Springer Nature B.V. 2021

Abstract
The graphitic carbon nitride quantum dots (CNQDs) and N-doped carbon quantum
dots (N-CQDs) could be prepared by the same raw material and synthetic method.
In this study, CNQDs and N-CQDs with the same QY (ca. 40%), functional group,
emission peak and similar particle size, but nitrogen functional group content, were
synthesized via a simple hydrothermal approach using ammonium citrate with and
without urea. Their different effects on the photodegradation property of T­ iO2 were
investigated via the synthesis of a series of T­ iO2/QDs samples through hydrother-
mal treatment, and the ­TiO2/QDs composite was comprehensively characterized.
Compared with pure ­TiO2, the photocatalytic efficiency of organic pollutants with
­TiO2 was enhanced by incorporation of QDs. Loading CNQDs on T ­ iO2 can improve
the generation of ·O2− radicals, and N-CQDs can increase the production of ·OH of
­TiO2. Therefore, the photocatalytic performance of the two composites to pollutants
is different.

Keywords  Graphitic carbon nitride dots · N-doped carbon dots · Photodegradation


efficiency

Introduction

The CNQDs and N-CQDs, the two zero-dimensional materials, have a wide vari-
ety of potential applications due to their unique electronic properties, biocompat-
ibility, tunable optical properties, high aqueous solubility and good stability [1, 2].
Both materials can be synthesized by two routes, namely, the top-down route and
* Bo Jin
jinbo0428@163.com
* Rufang Peng
rfpeng2006@163.com
1
State Key Laboratory of Environment‑Friendly Energy Materials, School of Materials Science
and Engineering, Southwest University of Science and Technology, Mianyang 621010, Sichuan,
People’s Republic of China

13
Vol.:(0123456789)
5230 P. Zhao et al.

the bottom-up route [3]. For the latter, CNQDs and N-CQDs are obtained using the
same raw material as a precursor. For example, high fluorescent CNQDs can be syn-
thesized using citric acid and urea as precursors through the microwave-assisted sol-
vothermal method [4], hydrothermal method [5], or solid-phase microwave-assisted
technique [6]. Meanwhile, water-soluble luminescent N-CQDs were also reported
using citric and urea as precursors through the microwave route [7], hydrothermal
method [8, 9] and solvothermal method [10]. CNQDs can be also obtained via a
solid-phase pyrolyzing melamine and ethylenediaminetetraacetic acid (EDTA) [11].
N-CQDs can also be synthesized through a one-step pyrolytic method using urea
and EDTA as precursors [12, 13]. CNQDs can be prepared by the polymerization of
­CCl4 as carbon source and 1,2-ethylenediamine as the nitrogen source under reflux,
microwave or solvothermal heating [14]. N-CQDs can also be obtained using 1,
2-ethylenediamine as the nitrogen source, and citric acid [15, 16], alanine [17], argi-
nine [18] or cysteine [19] as the carbon source. It can be postulated that the atomic
structure transforms from N-CQDs to CNQDs was caused by the substitutional N
gradually replaces pyrrolic or pyridinic N at a high nitrogen/carbon weight ratio [6].
Given their excellent properties, both QDs have been employed as efficient com-
ponents in the design of photocatalysts. Liu et al. [20] have reported that the separa-
tion of electron–hole pairs can be efficiently improved because the electrons in the
conduction band (CB) of g-C3N4 can easily transfer to N-CQDs. Li et al. [21] have
indicated that photocatalytic hydrogen evolution was enhanced by 11-fold compared
with that of pure g-C3N4 because the photogenerated electrons in the CB of g-C3N4
can transfer to the CB of CNQDs, and the holes in the valence band (VB) of CNQDs
can transfer to that of g-C3N4. N-CQDs can also enhance the photocatalytic activ-
ity of ­BiPO4 attributed to its molecular oxygen activation ability [22]. CNQDs can
enhance the photocatalytic activity for the degradation of orange photodegradation
of ­BiPO4 by creating a novel heterostructure [23]. Thus, CNQDs and N-CQDs are
not only synthesized by the same method and raw material by changing the nitro-
gen/carbon weight ratio, but they can also improve the photocatalytic performance
of semiconductors. However, the difference between the effects of CNQDs and
N-CQDs on semiconductors has not been reported thus far.
Among various photocatalysts, ­TiO2 remains the most used semiconductor pho-
tocatalyst for organic pollutant degradation [24], water splitting [25] and ­CO2 reduc-
tion [26] due to its chemical stability, excellent stability and low cost. Chen et  al.
[27] believed that the interfacial charge transfer and separation in heterojunction
formed by CNQDs and ­TiO2 were improved. Lin et  al. [28] indicated that n-type
­TiO2 can turn to a p-type one so that it can easily transfer electrons to CQDs after
loading CQDs. Thus, the photocatalytic effect of ­TiO2 can be improved effectively
by both QDs.
Herein, CNQDs and N-CQDs were synthesized by varying the urea ammonium
citrate to urea ratio, and different CNQDs/N-CQDs modified ­TiO2 composites were
prepared. The optical, textural, morphological and chemical properties of the as-
prepared samples were analyzed by photoluminescence (PL), transmission electron
microscopy (TEM), scanning electron microscopy (SEM) and X-ray diffraction
(XRD) and X-ray photoelectron spectroscopy (XPS) techniques. The photocatalytic
abilities were evaluated via the photodegradation of ciprofloxacin (CIP), tetracycline

13
Facile synthesis of quantum dots/TiO2 photocatalyst with… 5231

(TC), p-nitrophenol (PNP), 2,4-dinitrophenol (2,4-DNP), methylene blue (MB),


Solferino and Rhodamine B (RhB) as targeted contaminants. At the same time, the
effects of different parameters (pH and the initial of concentration of CIP) on pho-
toreaction were also investigated. Finally, the mechanism of photodegradation was
evaluated.

Experimental

Preparation of photocatalyst

CNQDs were prepared based on our previous work [29]. In details, urea and ammo-
nium citrate (mass ratio of 18:1) were dissolved in 2  ml water for and heated at
180 ℃ for 3 h. Then, a 0.22 μm membrane was employed to filter large particles.
The solution was added into 50 mL of deionized water and kept 2℃ for further use.
N-CQDs were synthesized by heating ammonium citrate solution without urea in
the same hydrothermal process.
The ­TiO2/QDs composites were synthesized via a one-step hydrothermal method:
in detail, 200 mg T
­ iO2 and amount of QDs solutions. After ultrasonic treatment for
30 min, the mixture was transferred into Teflon-lined stainless steel autoclave and
kept at 150 ℃ for 2 h. Then, the powder was collected by filtration and washed with
deionized water several times and dried in an oven at 50 ℃ overnight. For conveni-
ence, such obtained product was labeled as ­TiO2/QDs-x (x), where x represented the
volume of QDs added in the preparation process, and letter C represents CNQDs
and N represents N-CQDs.

Characterization

The morphologies of as-prepared samples were performed by SEM (Libra 200


PE), and TEM (Libra 200, Cari Zeiss Irts). The crystal structures were meas-
ured by a Philip X’ Pert PRO (Netherlands). The chemical states of element in
as-prepared samples were confirmed by XPS using a ThermoFisher system. The
Brunauer–Emmett–Teller (BET) surface areas were calculated by nitrogen adsorp-
tion–desorption isotherms using a quantachrome Autosorb-IQ surface area. Dif-
fuse reflectance spectra (DRS) were obtained on a Solidspec-3700 spectrophotom-
eter with ­BaSO4 as reference. The optical properties of aqueous were studied by
a Shimadzu F-6000 fluorospectro photometer and an Agilent Cary 60 UV–visible
spectrometer. The FT-IR spectra were recorded on a Nicolet 5700 spectrometer.
The transient photocurrent, electrochemical impedance spectroscopy (EIS) and
Mott–Schottky curves were performed on an electrochemical analyzer (CHI 760e
electrochemical workstation). The electron spin resonance (ESR) was carried out on
a Bruker EMX nano ESR spectrometer to detect reactive species with 5,5-dimethyl-
1-pyrroline-N-oxide (DMPO) as the radical’s spin-trapped reagent.

13
5232 P. Zhao et al.

Scheme 1  Schematic illustration of a CNQDs, and b N-CQDs preparation

Photocatalytic activity

The adsorption and photocatalytic activities of as-prepared photocatalysts were


tested for the degradation of organic pollutions under simulated sunlight irra-
diation at room temperature. In a typical procedure, 20  mg of the catalyst was
dispersed in 50 mL of organic pollutions aqueous solution and stirred for 0.5 h
in the dark to reach adsorption equilibrium before irradiation. Then, the sus-
pensions were illuminated with a 500 W Xe lamp. During the photodegradation
process, 4  mL suspension was extracted at predetermined time intervals, and
centrifugation to remove the photocatalyst. The remaining organic pollutions
concentration was determined by a UV–Vis spectrophotometer.

Results and discussion

Preparation of QDs

As shown in Scheme 1, CNQDs and N-CQDs can be synthesized by a precisely


controlled composition. Urea is a common nitrogen source for synthesizing
CNQDs and N-CQDs, and ammonium citrate is a frequent carbon source for
synthesizing N-CQDs [30, 31]. A large amount of O decoration (e.g., carboxylic
and hydroxyl) is present in the surface of CNQDs and N-CQDs, indicating the
high water solubility of both QDs.

13
Facile synthesis of quantum dots/TiO2 photocatalyst with… 5233

Fig. 1  Absorbance (red line), excitation and emission spectra of a CNQDs, and b N-CQDs

Fig. 2  TEM images and magnified HRTEM image (inset) of a CNQDs and b N-CQDs

Characterization of QDs

As shown in Fig. 1, both QDs show the two clear UV–Vis absorption bands at 234
and 340 nm, which can be ascribed to the presence of the π → π* and n → π* transi-
tions [32], respectively. And the aqueous solution of both QDs exhibits excitation-
wavelength-independent PL properties with emission peaks at 440 nm at excitation
from 320 to 420 nm. The relatively narrow half maximum to some extent confirm-
ing the narrow size distribution of the QDs [33].
The strongest fluorescence emission band of CNQDs and N-CQDs was observed
under 352 and 375 nm, respectively. The fluorescent quantum yield (QY) of CNQDs
and N-CQDs was 40.5% and 39.4%, respectively. Thus, we can rule out the influ-
ence of the QY of QDs on the photocatalytic performance of ­TiO2.
As shown in Fig. 2a and b, both of CNQDs and N-CQDs are uniformly mono-
dispersed, and the lattice spacings of CNQDs and N-CQDs are 0.31 and 0.18 nm,

13
5234 P. Zhao et al.

which correspond to the (002) plane of g-C3N4 [34], and the (102) facet of graphitic
carbon [35], respectively. As estimated from the TEM images, the diameters of
CNQDs and N-CQDs are mainly distributed in the narrow range of 2–6  nm, and
2–8 nm, respectively.
The FT-IR spectra of CNQDs and N-CQDs are shown in Fig.  3. For the two
QDs, the broad peak at 3000–3600 ­cm−1 corresponds to O–H and N–H stretching
vibrations [36]. The peaks located at 1605 and 1710 ­cm−1 belong to C=O stretch-
ing absorption [37]. The peak at 1397 ­cm−1 is the C=N vibration bonds [38]. The
results indicate that CNQDs and N-CQDs are composed of O–H, N–H, C = O and
C=N.
As shown in Fig. 4a, CNQDs and N-CQDs are composed of C, O and N. CNQDs
have more N content than N-CQDs, and N content is critical for photocatalysis [39].
As shown in Fig.  4b, the C 1  s high-resolution spectrum exhibits three peaks at
288.5, 286.4 and 284.8 eV, which are assigned to COOH, C-N and C–C/C=C [40],
respectively. The relative intensity of the C-N peak of CNQDs (18.9%) was slightly
higher than that of N-CQDs (16.5%). Figure 4c shows two peaks located at 401.5
and 399.7 eV in the high-resolution N, which are from the graphitic N and pyridinic
N, respectively. The areal ratio of pyridinic N to graphitic N in CNQDs was greater
than 1.4. confirming the presence of g-C3N4 [41]. Naturally, N-CQDs contain more
graphitic N than CNQDs. In Fig.  4d, the strong peaks at 533.3 and 531.5  eV can
be attributed to the C=O and C–OH bonds [42]. This result is in line with that of
FT-IR. Thus, the as-prepared CNQDs and N-CQDs have the same element, func-
tional group, emission peak and QY and have similar particle size, but their nitrogen
functional group content is different (Table. S1).

Characterization of ­TiO2/QDs composites

In Fig.  5a, the diffraction peaks with 2θ values at 25.28°, 37.8°, 48.0°, 53.89°,
55.06° and 62.68° can be indexed to the (101), (004), (200), (105), (211) and (204)

Fig. 3  FT-IR spectra of CNQDs


and N-CQDs

13
Facile synthesis of quantum dots/TiO2 photocatalyst with… 5235

Fig. 4   a XPS survey spectra, and High-resolution b C 1 s, c N 1 s and d O 1 s spectra of CNQDs and
N-CQDs

Fig. 5   a XRD pattern, and b FT-IR of photocatalysts

and depict the anatase phase of T


­ iO2 (PDF#01-071-1166). The as-prepared samples
show similar XRD patterns with the anatase ­TiO2, and no new diffraction peaks
appeared due to the low amount of QDs, indicating that QDs does not affect the

13
5236 P. Zhao et al.

crystal structure of ­TiO2. In Fig. 5b, the broad peak around 3434 ­cm−1 is attributed
to the absorption of water on the T ­ iO2 surface. The peak at 1633 ­cm−1 corresponds
to the stretching band of Ti–OH. The strong absorption at region 860 ­cm−1 corre-
sponds to the strong absorption of Ti–O–Ti [43]. Due to the small number and high
dispersion of QDs, there is no QD peak in the XRD and FT-IR spectra.
As shown in Fig.  6a, the survey spectrum of composites indicates the pres-
ence of Ti, C, N and O. The element N indicates the QDs load on the surface of
­TiO2 successfully. In Fig. 6b, the high-resolution Ti 2p data display two peaks at
458.1 and 463.7 eV ascribed to Ti 2­ p3/2 and Ti ­2p1/2, respectively, and the split-
ting of 5.6 eV designates the oxidation states of ­Ti4+ [44]. For the XPS spectrum
of the N 1  s signal (Fig.  6c), only one peak at 399.5  eV corresponds to C-N–C
in the T­ iO2/N-3.5 composites, while two peaks at 400.8 and 399.1  eV corre-
spond to N-C3 and C–N–C, respectively, in T ­ iO2/C-2.0 composites. The graphitic
N could facilitate the conductivity of T ­ iO2, and the π-electrons in the pyridine
ring could accelerate the separation of photogenerated carrier [45]. In Fig.  6d,
the higher peak is for the Ti–O-Ti linkage in T ­ iO2 at 529.3 eV, and the other one
is adsorbed water at 530.0 Ev [46]. The peaks of Ti-C bond (282.0 eV) and Ti-N

Fig. 6  XPS spectra of a full spectrum, High-resolution b Ti 2p spectra, c N 1 s spectra and d O 1 s spec-
tra of T
­ iO2/N-3.5 and T
­ iO2/C-2.0

13
Facile synthesis of quantum dots/TiO2 photocatalyst with… 5237

bond (396.0 eV) did not appear, indicating that C or N atoms are not doped into
the ­TiO2 lattice [47]. Compared with ­TiO2/C-2.0, the peaks of ­TiO2/N-3.5 shifted
to a higher binding energy, which indicates that the interaction between ­TiO2 and
N-CQDs is closer.
In Fig.  7a and b, ­TiO2 particles can be clearly observed, and the size of the
prepared sample is in the range of 100–400 nm. In Fig. 7c and d, QDs are highly
dispersed on the surface of T­ iO2 particles, indicating that the interaction between
­TiO2 and QDs is close, since the QDs on the surface of T ­ iO2 are not separated
by strong ultrasonication during the preparation process. The lattice spacing of
0.23 nm corresponds to the (004) plane of anatase ­TiO2.
As seen in Fig.  8, the specific surface areas of ­ TiO2/C-2.0 (9.395  ­m2/g)
2
and ­TiO2/N-3.5 (10.14  ­m /g) were slightly smaller than that of pure T ­ iO2
(16.977  ­m2/g), indicating that QDs were coated on the surface of ­TiO2 [48],
which could enlarge the particle size. Besides improving the separation of charge
carrier, the QDs could create new active sites onto the T ­ iO2 surface [49], then,
the photodegradation rate was enhanced markedly.

Fig. 7  SEM images of a ­TiO2/C-2.0 and b ­TiO2/N-3.5, and HRTEM images of c ­TiO2/C-2.0 and d
­TiO2/N-3.5

13
5238 P. Zhao et al.

Fig. 8  N2 adsorption–desorption
isotherms of photocatalysts

Photocatalytic performance

Loading QDs onto ­TiO2 dramatically promoted the photodegradation rate, and
the best samples were T ­ iO2/C-2.0 (Fig. S1a) and ­TiO2/N-3.5 (Fig. S1b). Exces-
sive amounts of QD will reduce the photocatalytic performance of ­TiO2 because
at high amounts, QDs might compete with T ­ iO2 for the photons [24]. The influ-
ence of the initial CIP concentrations (4, 8, 15 and 20 mg/L) on the photocatalytic
performance of the as-prepared samples was evaluated. In Fig.  9, as the initial
concentration of CIP increased, the degradation efficiency tended to decrease.
Then, CIP decomposed very slowly at 20 mg/L concentration. The photocatalytic
efficiency of ­TiO2/N-3.5 was slightly higher than that of ­TiO2/C-2.0.
Although the higher the initial concentration of CIP, the more the CIP mol-
ecules adsorbed on the surface of photocatalysis, free radicals formed on the
surface of ­TiO2 are the constant. Therefore, the relative number of free radicals
attacking the CIP molecules was decreased [50]. Meanwhile, a high concentra-
tion of CIP resulted in the formation more intermediates, which could occupy the
activity sites on the photocatalyst surface to reach a saturation point. Hence, the
photocatalytic efficiency decreased as the initial CIP concentration increased.
Based on the reaction kinetics shown in Fig.  10, which was fitted well into
the first-order kinetic equation – ln (C0/C) = k, the rate constant k of CIP photo-
degradation decreased from 0.0149  ­min−1 to 0.0038  ­min−1 for ­TiO2/C-2.0 and
from 0.0232 to 0.0048 ­min−1 for ­TiO2/N-3.5, as the initial concentration of CIP
increased from 4 to 20 mg/L. In the concentration range of 4–20 mg/L, the rate
constants of T ­ iO2/C-2.0 and T­ iO2/N-3.5 were approximately 2.3 and 1.4 times
and 3.5 and 1.7 times higher, respectively, than that of pure ­TiO2 (Fig. S2).
The descent scope of the rate constant of T ­ iO2/N-3.5 was greater than that of
­TiO2/C-2.0. We can assume that the ­TiO2/N-3.5 sample was sensitive to the ini-
tial concentration of the organic pollutant.

13
Facile synthesis of quantum dots/TiO2 photocatalyst with… 5239

Fig. 9  Effect of the initial concentrations a 4 mg/L, b 8 mg/L, c 15 mg/L and d 20 mg/L of CIP photo-
gradation under simulated sunlight with photocatalysts

Fig. 10  Corresponding reaction kinetic by a ­TiO2/C-2.0 and b ­TiO2/N-3.5 composites at different initial


CIP concentrations

These hypotheses were tested by investigating the effects of the initial con-
centration of RhB on the photocatalytic degradation of the as-prepared sample.
As  expected, the ­TiO2/N-3.5 sample showed the best photocatalytic activity at
RhB concentration of 20  mg/L (Fig. S3a). However, the activity declined when

13
5240 P. Zhao et al.

the RhB concentration was 30 mg/L (Fig. S3b). Meanwhile, the T ­ iO2/C-2.0 sam-
ple showed a better photocatalytic activity than the pure ­TiO2.
The photocatalytic performance of the as-prepared sample was evaluated by
examining the photodegradation of other organic pollutants. The T ­ iO2/N-3.5
sample still showed the highest solferino effective degradation (Fig. S4), while
the ­TiO2/C-2.0 catalyst exhibited the best photocatalytic degradation of MB, TC,
PNP and 2,4-DNP (Fig. S5). The degradation efficiency of the as-prepared sam-
ples for different pollutants is summarized in Table S1.
The effects of pH value on CIP degradation efficiency were tested by perform-
ing a set of experiments at different pH values. Because the charges on the sur-
face of the catalysts changed under different pH conditions, the photocatalysts
exhibited different absorptive capacities. Fig. S6 showed that acid solution (pH
2.0) decreased the amount of CIP molecule absorbed by ­TiO2, while acid (pH
2.0) and alkaline solutions (pH 13.0) could decrease the amount of CIP absorbed
by both as-prepared samples (Fig. 11).
The highest removal efficiency of T ­ iO2, ­TiO2/C-2.0 and ­ TiO2/N-3.5 was
observed at pH 13.0, 6.8 and 6.8, respectively, and the lowest removal efficiency
of ­TiO2, ­TiO2/C-2.0 and ­TiO2/N-3.5 was observed at pH 2.0, 13.0 and 2.0,
respectively, suggesting that the ­TiO2 surface was changed after it was coating
with QDs. The acid solution had a minimal effect on the removal efficiency of
­TiO2/C-2.0 but had a considerable influence on that of T ­ iO2/N-3.5.
The stability of a catalyst is another important factor for practical applica-
tions [51]. Figure 12 shows the stability of the composites after three consecutive
experiments for CIP degradation under simulated solar light irradiation. Com-
pared with the first cycle (78%), the photocatalytic efficiency of ­TiO2/C-2.0 in the
third cycle shows slight loss (67%), while for ­TiO2/N-3.5, the photocatalytic effi-
ciency almost remains the same after three cycles of experiments, which showed
the ­TiO2/N-3.5 had good reusable in photocatalytic activity. There is no distinct
change in the XRD patterns of the as-prepared samples after three cycles (Fig.
S7).

Fig. 11  Effect of initial pH on CIP photodegradation by a ­TiO2/C-2.0 and b ­TiO2/N-3.5

13
Facile synthesis of quantum dots/TiO2 photocatalyst with… 5241

Fig. 12  Recycling runs of the degradation of CIP over a ­TiO2/C-2.0, and b ­TiO2/N-3.5

Mechanism of photocatalytic analysis

The optical absorption of photocatalysts was examined by DRS spectroscopy. As


shown in Fig. 13a, the spectra of pure T­ iO2 show a typical absorption band in the
ultraviolet range, and the as-prepared samples display a small red shift compared
with the pure ­TiO2. The band gaps of ­TiO2, ­TiO2/C-2.0 and ­TiO2/N-3.5 had been
calculated as 3.35, 3.31 and 3.29 eV, respectively. To obtain the effect of the light
wavelength on the photocatalysts more accurately, SPV measurements were used
in this experiment. As shown in Fig. 13b, all samples generate distinguished SPV
responses under illumination. Apparently, the composites have a wilder photore-
sponse, and stronger SPV response signal, that is, better photogenerated elec-
tron–hole separation efficiency [52].
Transient photocurrent response and electrochemical impedance spectroscopy
(EIS) were tested to further verify the separation and recombination of carriers of
the as-prepared samples. In general, the high photocurrent response means that
photogenerated charges can be effectively separated and have high photocatalytic

Fig. 13  DRS spectra and SPV spectra of photocatalysts

13
5242 P. Zhao et al.

Fig. 14  a Photocurrent curves and b Nyquist plots of photocatalysts

Fig. 15  Mott–Schottky plots of a ­TiO2/C-2.0, and b ­TiO2/N-3.5

activity [53]. As shown in Fig. 14a, the current signal immediately appeared when
the light was on, and T ­ iO2/C-2.0 and T ­ iO2/N-3.5 presented drastically enhanced
current density compared with T ­ iO2, which demonstrated that QDs can effectively
facilitated the charge separation efficiency of ­TiO2 [54]. This result suggests a pos-
sible interaction between QDs and ­TiO2.
According to the Nyquist circle principle, a small arc diameter indicates high
conductivity, which is favorable to the transfer of photon-generated carriers [55].
Figure 14 (b) shows that the large arc of T­ iO2 in the EIS Nyquist plots confirmed the
highest rate of electron-transfer resistance [56]. This finding confirms that QDs can
lower the interfacial layer resistance of T ­ iO2, and ­TiO2/C-2.0 has highest electron
mobility.
In general, the CB of the samples was approximately determined by evaluating
the flat band potentials ­(Vfb) from Mott–Schottky measurement at 500, 750 and
1000 Hz. A positive slope of linear curve (1/C2 vs. potential) represents an n-type
one [57], indicating that ­TiO2 is an n-type semiconductor (Fig. S8). The linear
plots of T­ iO2/C-2.0 (Fig. 15a) and T­ iO2/N-3.5 (Fig. 15b) also show positive slopes,

13
Facile synthesis of quantum dots/TiO2 photocatalyst with… 5243

demonstrating that CNQDs and N-CQDs did not change the type of ­TiO2, and the
majority of charge carriers were still dependent on the Ti interstitials and O vacan-
cies [58]. The ­Vfb of T ­ iO2, ­TiO2/C-2.0 and ­TiO2/N-3.5 were approximately – 0.55,
– 0.46 and – 0.41 V, respectively (vs. Ag/AgCl). The calculated Vfb values of ­TiO2,
­TiO2/C-2.0 and ­TiO2/N-3.5 corrected by AgCl (vs. 0.22 eV Normal Hydrogen Elec-
trode (NHE)) were – 0.33, – 0.24 and – 0.19 V (vs. NHE) [59], respectively, indi-
cating that the Vfb position of ­TiO2 can be more positive due to QDs. For n-type
semiconductors, CB locates more negatively by 0.1–0.2  eV than the ­Vfb position.
Given that the voltage difference between CB and Vfb was 0.15 eV, the CB positions
of ­TiO2, ­TiO2/C-2.0 and ­TiO2/N-3.5 were – 0.48, – 0.39 and – 0.34 V (vs. NHE),
respectively, which are all lower than the standard O ­ 2/·O2− potential (– 0.33 V vs.
NHE) [60]. As such, all photocatalysts can reduce ­O2 to ·O2−. The band gaps of
­TiO2, ­TiO2/C-2.0 and T ­ iO2/N-3.5 were 3.35, 3.31 and 3.29  eV, respectively. The
VBs of T ­ iO2, ­TiO2/C-2.0 and T ­ iO2/N-3.5 were 2.87, 2.92 and 2.95 eV, respectively,
and high enough to oxidize H ­ 2O to produce ·OH potential (2.4 V, vs. NHE) [61].
The influences of the active species in CIP degradation by the as-prepared sam-
ple were confirmed by using potassium iodide (KI), benzoquinone (BQ) and iso-
propanol (IPA) to scavenge the active species h­ +, ·O2−, and ·OH, respectively [62].
As shown in Fig.  16, the degradation deficiency of both samples decreased obvi-
ously after BQ was added, suggesting that ·O2− reactive species had a major effect
on removing CIP. In addition, the degradation efficiency was similar after KI and
IPA were added, indicating the equally important role of ­h+ and ·OH reactive spe-
cies during the photocatalytic degradation of CIP. These results confirm that ·O2−
radicals are the primary active species in the system, while ­h+ and ·OH are also
involved. The results are consistent with those of ­TiO2 (Fig. S9).
The ESR spin-trapping technique (with DMPO) was used to study the effects
of the as-prepared samples on the formation of ·O2− and ·OH radicals under solar
light irradiation. As shown in Fig. 17a and d, for ­TiO2, the characteristic signal
of the DMPO- ·O2− and DMPO- ·OH radicals in obvious peak was detected when
the light was on, whereas no signals were observed in the dark. Thus, ·O2− and
·OH radicals were efficiently generated under simulated sunlight. Among them,

Fig. 16  Photocatalytic degradation of CIP after addition of various scavenging agents using a ­TiO2/C-2.0
and b ­TiO2/N-3.5 under simulated solar light irradiation

13
5244 P. Zhao et al.

Fig. 17  ESR spectra recorded ambient temperature in methanol dispersion with a ­TiO2, b ­TiO2/C-2.0
and c ­TiO2/N-3.5 and in aqueous dispersion with d ­TiO2, e ­TiO2/C-2.0 and f ­TiO2/N-3.5

the corresponding intensities of DMPO- ·O2− signals were obviously weakened


after 5  min illumination, while the ­ TiO2/C-2.0 displayed stronger signals of
DMPO- ·O2− (Fig. 17b), and T ­ iO2/N-3.5 did not show any difference (Fig. 17c).
Besides, the peak intensity of both composites was slightly changed with pro-
longed irradiation time. The results show that CNQDs can increase the generation
of ·O2− radicals of ­TiO2, and both QDs can maintain the yields of ·O2− radicals.
In contrast with that of T ­ iO2 (Fig.  17d), the DMPO- ·OH signal intensity of
­TiO2/C-2.0 was slightly increased under light irradiation (Fig. 17e), while the sig-
nal of ­TiO2/N-3.5 increased significantly (Fig. 17f), demonstrating the formation
of more ·OH radicals after simulated illumination for 2 min.
The above analysis is consistent with the result of ESR and free radical com-
petitive trapping experiments that both ·OH and ·O2− dominate the photodegrada-
tion of organics.
Based on the above results, the proposed energy band alignments and charge
transfer mechanism of T ­ iO2/C-2.0 and ­TiO2/N-3.5 sample are proposed. As illus-
trated in Fig.  18a, the highly active ·OH and ·O2− radical species confirm the
Z-scheme structure formed by ­TiO2 and CNQDs instead of the typical Type II
heterojunction structure [63, 64]. ­TiO2 and CNQDs were excited by light, and
the electrons in the CB of CNQDs reacted with ­O2 to form ·O2− [65], and holes
in the VB of T­ iO2 were trapped by adsorbed H ­ 2O to generate ·OH radicals, which
help oxidize organic pollutant [66]. The CB potential of CNQDs was more neg-
ative than the standard O ­ 2/·O2− potential, resulting in more generation of ·O2−.
In Fig.  18b, the electrons of T ­ iO2 excited form VB to the CB under simulated
sunlight, leaving the holes on VB, making it is the easier to produce ·OH, and
CQDs can act as electron containers as well as mediators to abstract the electrons

13
Facile synthesis of quantum dots/TiO2 photocatalyst with… 5245

Fig. 18  Schematic diagram for the photodegradation mechanism of a ­TiO2/C-2.0 and b ­TiO2/N-3.5 sam-
ple

emitted form T
­ iO2 [67, 68], due to their excellent electronic conductivity, sup-
pressing undesirable recombination process [69].

Conclusions

In conclusion, CNQDs and N-CQDs with the same QY (ca. 40%), functional group,
emission peak and similar particle size, but different nitrogen functional group con-
tent, were synthesized via a simple approach using ammonium citrate with and with-
out urea. Moreover, the QDs/TiO2 composites were successfully prepared by a facile
hydrothermal way and used for wastewater remediation. Incorporation of CNQDs
could improve the photocatalytic performance of T ­ iO2, since loading CNQDs can
improve the generation of ·O2− radicals of ­TiO2. Thus, ­TiO2/C-2.0 catalysts exhib-
ited excellent photocatalytic degradation performance for MB, TC, p-nitrophenol,
2,4-dinitrophenol and RhB (30  mg/L). N-CQDs can increase the ·OH production
of ­TiO2, so the T
­ iO2/N-3.5 sample showed the highest effective degradation perfor-
mance for CIP, solferino and RhB (20 mg/L).

Supplementary Information  The online version contains supplementary material available at https://​doi.​
org/​10.​1007/​s11164-​021-​04595-4.

Acknowledgements  This work was carried out with the financial support received from the Natural Sci-
ence Foundation of China (21875192), Outstanding Youth Science and Technology Talents Program
of Sichuan (19JCQN0085), and Open Project of State Key Laboratory of Environment-friendly Energy
Materials (20fksy16).

References
1. H. Liu, X. Wang, H. Wang, R. Nie, J. Mater. Chem. B. 7, 5432 (2019)
2. S.Y. Lim, W. Shen, Z.Q. Gao, Chem. Soc. Rev. 44, 362 (2015)
3. T. Wang, C. Nie, Z. Ao, S. Wang, T. An, J. Mater. Chem. A. 8, 485 (2020)
4. X.T. Cao, J. Ma, Y.P. Lin, B.X. Yao, F.M. Li, Spectrochim. Acta A 151, 875 (2015)

13
5246 P. Zhao et al.

5. G.Y. Bai, Z.P. Song, H.Y. Geng, D. Gao, K. Liu, S.W. Wu, W. Rao, L.Q. Guo, J.J. Wang, Adv.
Mater. 29(28), 1606843 (2017)
6. S.Y. Gu, C.T. Hsieh, Y.A. Gandomi, J.K. Chang, J. Li, J.L. Li, H. Zhang, Q. Guo, K.C. Lauh, R.
Pandey, J. Mater. Chem. C. 7, 5468 (2019)
7. L. Sciortino, A. Sciotino, R. Popescu, R. Schneider, D. Gerthsen, S. Agnello, M. Cannas, F.
Messina, J. Phys. Chem. C. 122(34), 19897 (2018)
8. B.B. Wang, Y.F. Wang, H. Wu, X.J. Song, X. Guo, D.M. Zhang, X.J. Ma, M.Q. Tan, RSC Adv. 4,
49960 (2014)
9. Y. Zhang, Y.H. He, P.P. Cui, X.Y. Feng, L. Chen, Y.Z. Yang, RSC Adv. 5, 40393 (2015)
10. K. Hola, M. Sodolska, S. Katlytchuk, D. Nachtigallová, A.L. Rogach, M. Otyepka, R. Zbořil, ACS
Nano 11(12), 12402 (2017)
11. X.Q. Fan, Y. Feng, Y.Y. Su, L.C. Zhang, Y. Lv, RSC Adv. 5, 55158 (2015)
12. K. Patir, S.K. Gogoi, Nanoscale Adv. 1, 592 (2019)
13. C.P. Han, R. Wang, K.Y. Wang, H.Y. Xu, M.R. Sui, Biosens. Bioelectron. 83, 229 (2016)
14. S. Jiu, J.Q. Tian, L. Wang, Y.L. Luo, J.F. Zhai, X.P. Sun, J. Mater. Chem. C. 21, 11726 (2011)
15. S.J. Zhu, Q.N. Meng, L. Wang, J.H. Zhang, Y.B. Song, H. Jin, K. Zhang, H.C. Sun, H.Y. Wang, B.
Yang, Angew. Chem., Int. Ed. 52(14), 3953 (2013)
16. Y.L. Wang, J.X. Zheng, J.L. Wang, Y.Z. Yang, X.G. Liu, Opt. Mater. 73, 319–329 (2017)
17. W.J. Niu, Y. Li, R.H. Zhu, D. Shan, Y.R. Fan, X.J. Zhang, Sensor. Actuat. B: Chem. 218(31), 229
(2015)
18. F. Arcudi, L. Dordevic, M. Prato, Angew. Chem., Int. Ed. 128(6), 2147 (2016)
19. S.F. Xu, S.Q. Ye, Y.H. Xu, F.F. Liu, Anal. Sci. 36(3), 353 (2020)
20. F.L. Wang, P. Chen, Y.P. Feng, Z.J. Xie, Y. Liu, Y.H. Su, Q.X. Zhang, Y.F. Wang, K. Yao, W.Y. Lv,
G.G. Liu, Appl. Catal. B: Environ. 207, 103 (2017)
21. Y.P. Wang, Y.K. Li, J.L. Zhao, J.S. Wang, Z.J. Li, Int. J. Hydrogen Energ. 44, 618 (2019)
22. J. Di, J.X. Xia, X.L. Chen, M.X. Ji, S. Yin, Q. Zhang, H.M. Li, Carbon 114, 601–607 (2017)
23. Z.S. Li, B.L. Li, S.H. Peng, D.H. Li, S.Y. Yang, Y.P. Fang, RSC Adv. 4, 35144 (2014)
24. Q. Wang, J.S. Cai, G.V. Biesold-McGee, J.Y. Huang, Y.H. Ng, H.T. Sun, J.P. Wang, Y.K. Lai, Z.Q.
Lin, Nano Energy 78, 105313 (2020)
25. M.S. Nasirae, G.R. Yang, I. Ayub, S.L. Wang, W. Yan, Appl. Catal. B: Environ. 270, 118900 (2020)
26. M.E. Aguirre, R.X. Zhou, A.J. Eugene, M.I. Guzman, M.A. Grela, Appl. Catal. B: Environ. 217,
485 (2017)
27. J.Y. Su, L. Zhu, P. Geng, G.H. Chen, J. Hazard. Mater. 316, 159 (2016)
28. J.T. Zhang, Y.H. Cao, P. Zhao, T.F. Xie, Y.H. Lin, Z. Mu, Colloid Surf. A 601, 125019 (2020)
29. P. Zhao, B. Ji, Q.C. Zhang, R.F. Peng, Langmuir 37(5), 1760 (2021)
30. Y.J. Zhang, R.R. Yuan, M.L. He, G.C. Hu, J.T. Jiang, T. Xu, L. Zhou, W. Chen, W.D. Xiang, X.J.
Liang, Nanoscale 9, 17849 (2017)
31. Y.W. Hu, Y.X. Wang, C.T. Wang, Y.W. Ye, H.C. Zhao, J.L. Li, X.J. Lu, C.L. Mao, S.J. Chen, J.M.
Mao, L.P. Wang, Q.J. Xue, Carbon 152, 511 (2019)
32. W.F. Chen, D.J. Li, L. Tian, W. Xiang, T.Y. Wang, W.M. Hu, Y.L. Hu, S.N. Chen, J.F. Chen, Z.X.
Dai, Green Chem. 20, 4438 (2018)
33. X.M. Li, J.L. Chang, F. Xu, X.R. Wang, Y.H. Lang, Z.Y. Gao, D.P. Wu, K. Jiang, Res. Chem.
Intermed. 41, 813 (2015)
34. S. Zhang, J. Li, M. Zeng, J. Xu, X. Wang, W. Hu, Nanoscale 6, 4157 (2014)
35. S. Sahu, B. Behera, T.K. Maiti, S. Mohapatra, Chem. Commun. 48(70), 8835 (2012)
36. S.N. Qu, X.Y. Wang, Q.P. Lu, X.Y. Liu, L.J. Wang, Chem. Int. Ed. 51, 12215 (2012)
37. Y.Y. Hu, R.T. Guan, C. Zhang, K.Y. Zhang, W.J. Liu, X.D. Shao, Q.W. Xue, Q.L. Yue, Appl. Surf.
Sci. 531, 147344 (2020)
38. H. Li, F.Q. Shao, H. Huang, J.J. Feng, A.J. Wang, Sensor. Actuat. B: Chem. 226, 506 (2016)
39. A. Kumar, P. Raizada, A.H. Bandegharaei, V.K. Thakur, V.H. Nguyen, P. Singh, J. Mater. Chem. A.
9, 111 (2021)
40. Y. Lu, J. Chen, A.J. Wang, N. Bao, J.J. Feng, W.P. Wang, L.X. Shao, J. Mater. Chem. C. 3, 73
(2015)
41. S.Y. Gu, C. Te Hsieh, Y.A. Gandomi, J.L. Li, X.X. Yue, J.K. Chang, Nanoscale 11, 16553–16561
(2019)
42. L. Zhou, Y.H. Tian, J.Y. Lei, L.Z. Wang, Y.D. Liu, J.L. Zhang, Catal. Sci. Technol. 8, 2617 (2018)
43. J.Y. Su, L. Zhu, G.H. Chen, Appl. Catal. B: Environ. 186, 127 (2016)
44. M.S. Nasir, G.R. Yang, I. Ayub, S.L. Wang, W. Yan, Appl. Surf. Sci. 519, 146208 (2020)

13
Facile synthesis of quantum dots/TiO2 photocatalyst with… 5247

4 5. X. Xu, L. Lai, T. Zeng, Y. Yub, Z.Q, He, J. M Chen, S. Song. 122, 33(2018).
46. T.S. Zhou, S. Chen, L.S. Li, J.C. Wang, Y. Zhang, J.H. Li, J. Bai, L.G. Xia, Q.J. Xu, M. Rahim, B.X.
Zhou, Appl. Catal. B: Environ. 269, 118776 (2020)
47. Y.Y. Chen, Q.M. Dong, L.L. Wang, X. Guo, S.S. Ai, H.M. Ding, Res. Chem. Intermed. 44, 7369
(2018)
48. S. Shafafi, A.H. Yangjeh, S. Feizpoor, S. Ghosh, T. Maiyalagan, Sep. Purif. Technol. 250, 117179
(2020)
49. P. Singh, P. Shandilya, P. Raizada, A. Sudhaik, A.R. Sani, A.H. Bandegharaei, Arab. J. Chem. 13,
3498 (2020)
50. J.W. Zhang, D.F. Fu, Y.D. Xu, C.Y. Liu, J. Environ. Sci. 22(8), 1281 (2010)
51. J. Yu, J.Y. Lei, L.Z. Wang, C. Guillard, J.L. Zhang, Y.D. Liu, M. Anpo, Res. Chem. Intermed. 45,
4237 (2019)
52. Y.Y. Zhu, W. Deng, L. Chen, J. Courtois, Q. Tian, X.W. Zhang, L. Almasy, M.H. Yan, K. Xiong,
Appl. Surf. Sci. 534, 147646 (2020)
53. W.D. Dai, L. Jiang, J. Wang, Y.J. Pu, Y.F. Zhu, Y.X. Wang, B.B. Xiao, Chem. Eng. J. 397, 125476
(2020)
54. X.C. Yang, Y.L. Yang, S.L. Zhang, Y.F. Liu, S.J. Fu, M. Zhu, J.F. Hu, Z.J. Zhang, J.T. Zhao, Appl.
Surf. Sci. 490, 592 (2019)
55. X. Han, L. An, Y. Hu, Y.G. Li, C.Y. Hou, H.Z. Wang, Q.H. Zhang, Appl. Catal. B: Environ. 265,
118539 (2020)
56. P.K. Raizada, A. Sudhaik, P. Singh, A.H. Bandegharaei, P. Thakur, Sep. Purif. Technol. 227, 115692
(2019)
57. J. Yang, H. Miao, W.L. Li, H.Q. Li, Y.F. Zhu, J. Mater. Chem. A. 7, 6482 (2019)
58. R. Adzhri, M.K.M. Arshad, S.C.B. Gopinath, A.R. Ruslinda, M.F.M. Fathil, C. Ibau, M.N.

Nuzaihan, Sensor. Actuat. A:Phys. 259, 57–67 (2017)
59. G.P. Smestad, S. Spiekermann, J. Kowalik, C.D. Grant, A.M. Schwartzberg, J. Zhang, L.M. Tolbert,
E. Moons, Sol. Energ. Mat. Sol. C. 76(1), 85 (2003)
60. W.D. Dai, L. Jiang, J. Wang, Y.J. Pu, Y.F. Zhu, Y.X. Wang, B.B. Xiao, Chem. Eng. J. 397, 125476
(2020)
61. Y.H. Li, K.L. Lv, W.K. Ho, F. Dong, X.F. Wu, Y. Xia, Appl. Catal. B: Environ. 202, 611 (2017)
62. J. Yan, B. Jin, P. Zhao, R.F. Peng, Inorg. Chem. Front. 8, 777 (2021)
63. J.Q. Yan, H. Wu, H. Chen, Y.X. Zhang, F.X. Zhang, S.F. Liu, Appl. Catal. B: Environ. 191, 130
(2016)
64. L. Kong, X.T. Zhang, C.H. Wang, J.P. Xu, X.W. Du, L. Li, Appl. Surf. Sci. 488, 288 (2018)
65. P. Raizada, A. Sudhaik, P. Singh, P. Shandilya, V.K. Gupta, A.H. Bandegharaei, S. Agrawal, J. Pho-
toch. Photobio. A. 374, 22 (2019)
66. S.V. Dutta, S. Sharma, P. Raizada, A.H. Bandegharaei, V.K. Gupta, P. Singh, J. Saudi Chem. Soc.
23(8), 1119 (2019)
67. R.D. Liu, H. Li, L.B. Duan, H. Shen, Y.Y. Zhang, X.R. Zhao, Ceram. Int. 43(12), 8648 (2017)
68. V. Hasija, A. Sudhaik, P. Raizada, A.H. Bandegharaei, P. Singh, J. Environ Chem. Eng. 7, 103272
(2019)
69. L. Midya, A.N. Sarkar, R. Das, A. Maity, S. Pal, Int. J. Biol. Macromol. 164, 3676–3686 (2020)

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published
maps and institutional affiliations.

13

You might also like