You are on page 1of 47

Journal Pre-proofs

Enhanced photocatalytic degradation of antibiotics in water over functionalized


N,S-doped carbon quantum dots embedded ZnO nanoflowers under sunlight
irradiation

Yanning Qu, Xiaojian Xu, Renliang Huang, Wei Qi, Rongxin Su, Zhimin He

PII: S1385-8947(19)32426-X
DOI: https://doi.org/10.1016/j.cej.2019.123016
Reference: CEJ 123016

To appear in: Chemical Engineering Journal

Received Date: 8 July 2019


Revised Date: 26 September 2019
Accepted Date: 29 September 2019

Please cite this article as: Y. Qu, X. Xu, R. Huang, W. Qi, R. Su, Z. He, Enhanced photocatalytic degradation of
antibiotics in water over functionalized N,S-doped carbon quantum dots embedded ZnO nanoflowers under sunlight
irradiation, Chemical Engineering Journal (2019), doi: https://doi.org/10.1016/j.cej.2019.123016

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.

© 2019 Published by Elsevier B.V.


Enhanced photocatalytic degradation of antibiotics in water over

functionalized N,S-doped carbon quantum dots embedded ZnO

nanoflowers under sunlight irradiation

a, 1 a, 1 b,
Yanning Qu , Xiaojian Xu , Renliang Huang *, Wei Qi a, c, d, *, Rongxin Su a, c, d
,

Zhimin He a

a
State Key Laboratory of Chemical Engineering, School of Chemical Engineering and

Technology, Tianjin University, Tianjin 300072, P. R. China

b
Tianjin Key Laboratory of Indoor Air Environmental Quality Control, School of

Environmental Science and Engineering, Tianjin University, Tianjin 300072, P. R. China

c
Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin

300072, P. R. China

d
Tianjin Key Laboratory of Membrane Science and Desalination Technology, Tianjin

University, Tianjin 300072, P. R. China

1
These two authors contributed equally to this work.

* Author to whom any correspondence should be addressed

E-mail: tjuhrl@tju.edu.cn (R. H.); qiwei@tju.edu.cn (W. Q.)

Tel: +86 22 27407799. Fax: +86 22 27407599.


Abstract: ZnO is widely used as a photocatalyst in industry, however, it is still a challenge
to degrade refractory antibiotics in water. In this study, a novel surface-functionalized

N,S-doped carbon quantum dot (N,S-CQD) was synthesized and embedded into ZnO for the

formation of a new ZnO/N,S-CQDs hybrid nanoflower via one-pot hydrothermal process.

The as-prepared ZnO/N,S-CQDs showed significantly enhanced photocatalytic activity under

visible and near-infrared (NIR) light irradiation, in which 72.8% of MG was decomposed

after 180 min under NIR light. In addition, approximately 92.9% and 85.8% of ciprofloxacin

(CIP) were degraded by ZnO/N,S-CQDs under simulated sunlight for 20 min and natural

sunlight for 50 min, respectively. Furthermore, the mechanism was investigated and the

results show that the surface functionalization, electron transfer, up-converted luminescence

properties of N,S-CQDs, together with the highly reactive facets of ZnO nanoflowers, made

great contributions to the enhanced photocatalytic activity of ZnO/N,S-CQDs. Additionally,

the results of active species trapping experiments indicated that the hydroxide free radicals,

holes and superoxide radical anions all played certain roles in the photocatalytic reaction.

Finally, ZnO/N,S-CQDs was employed for photodegradation of antibiotics in actual water,

the degradation efficiency of antibiotics still remained above 60% after 120 min. We believe

that the ZnO/N,S-CQDs nanoflower is a promising photocatalyst for the degradation of

refractory antibiotics under sunlight irradiation. The relatively low cost and excellent

photocatalytic performance of ZnO/N,S-CQDs is beneficial for industrial applications.

Keywords: antibiotics; carbon quantum dots; ZnO nanoflowers; photocatalysis; sunlight

irradiation
1. Introduction
Antibiotics, a class of organic molecules, are often used for the prevention and therapy of

diseases concerning bacterial infections. With increasing overuse and misuse of antibiotics,

their global emissions have gradually increased in recent years [1, 2]. For example, the data

from a recent study showed that approximately 92,700 tons of antibiotics were consumed in

China in 2013, approximately 46% of which were released into rivers and lands, resulting in

serious environmental pollution [3]. Surprisingly, some antibiotics were even detected in

drinking water in many countries [4, 5], which is harmful to human health due to not only the

generation of drug resistance but also the breakdown of anthropic therapeutic potentials

against more serious diseases [6]. Therefore, it is highly desirable to develop environmentally

friendly and highly efficient methods for the disposal of antibiotics.

To date, many kinds of photocatalysts have been used for the photodegradation of

antibiotics. For example, Table S1 summarizes some previous reports concerning the

photocatalytic degradation of ciprofloxacin (CIP) under simulated sunlight. In general, in

comparison with traditional photocatalysts (such as TiO2 [7-9], ZnO [10] and γ-Fe2O3 [11]),

the new-typed photocatalysts (such as g-C3N4 [6, 12], BiOCl [13] and Bi2MoO6 [14]) showed

higher absorption of light in the visible region and exhibited better photocatalytic

performance. However, these traditional photocatalysts have been commonly used in industry

owing to their low cost and facile large-scale synthesis. Nevertheless, their photocatalytic

efficiencies were often restricted due to their broad band gaps for the photodegradation of

pollutants under full spectrum of sunlight, especially for some refractory antibiotics [8, 10,

15]. Therefore, the main challenge is to improve the photocatalytic performance of the
traditional photocatalysts so that they were more effective for the industrial applications.

As one of the effective approaches for improving photocatalytic efficiency, quantum dot

sensitization was generally taken into consideration. To date, this method has been used in

many fields of photoelectrocatalysis, such as solar cells [16], hydrogen production [17] and

pollutant treatment [18], in which the photocatalytic activities of the hybrid materials were

significantly improved via the assistance of quantum dot. Among these quantum dots, carbon

quantum dots (CQDs) are a promising candidate due to their high chemical stability,

photoinduced electron transfer and especially the excellent upconversion luminescence

properties [19, 20]. For example, Tian et al. [21] constructed a CQDs/H-TiO2 nanobelt

heterostructure for the photodegradation of methyl orange, achieving a degradation efficiency

of 50% after 25 min under visible light. Wang et al. [22] employed CQDs as a sensitizer and

synthesized a ternary SDAg-CQDs/UCN (ultrathin g-C3N4) catalyst, which showed 87.5%

degradation efficiency for naproxen under visible light, while the value for the SDAg/UCN

system was only 58.9%. In addition, graphene quantum dots (GQDs) were also used for the

synthesis of a composite GQDs/Mn-N-TiO2/g-C3N4 photocatalyst, in which 89% of

p-nitrophenol was degraded after 120 min under simulated sunlight [17]. In these studies, the

fundamental photocatalysts played important roles before the decoration of CQDs, i.e., their

photocatalytic efficiencies mainly depended on the basal materials rather than the CQDs.

Therefore, the key issue is how to design new CQDs matched with the primary photocatalyst,

making the CQD-embedded composite to be more simplified, high-efficiency and harmless

during the photocatalytic degradation of environmental pollutants. Additionally, the

post-modification of photocatalyst (e.g., ZnO) with quantum dots is generally employed and
it is a key issue in reducing the loss of quantum dots for recycling and reuse.

In this study, we designed and synthesized a new kind of surface-functionalized N,S-doped

CQDs (N,S-CQDs) via the hydrothermal treatment by using L-cysteine as a carbon source

and ethylene glycol as a passivation agent. Furthermore, the N,S-CQDs were embedded in

the ZnO nanoflowers with highly reactive facets via a one-pot hydrothermal process, leading

to the controllable synthesis of a three-dimensional (3D) heterostructured nanoflower

(denoted as ZnO/N,S-CQDs, Scheme 1). The structure of the composite was characterized,

and the photocatalytic performance was evaluated under visible light, near-infrared (NIR)

light, simulated sunlight, and natural sunlight irradiation by using CIP, cephalexin (CEL),

methylene blue (MB), rhodamine B (RhB) or malachite green (MG) as model pollutants. To

illustrate the photocatalytic mechanism, three other CQDs were synthesized for comparison,

in which one is the N,S-doped CQDs without surface modification, the other two is

Try-CQDs and Glu-CQDs via hydrothermal treatment of L-tryptophan and L-glutamic acid,

respectively. In addition, the active species trapping experiments, the photocurrent responses

of hybrid nanoflowers and the up-converted photoluminescence property of N,S-CQDs were

also performed. Finally, the photocatalytic degradation of antibiotics (CIP, CEL) in actual

water by the novel ZnO/N,S-CQDs photocatalyst were also tested under simulated sunlight to

evaluate its practical use in water treatment.


2. Experimental
2.1. Materials

Zinc nitrate hexahydrate (Zn(NO3)2·6H2O), ethylene glycol (EG), sodium hydroxide

(NaOH) and the organic pollutants methylene blue (MB), rhodamine B (RhB), and malachite

green (MG) were obtained from Guangfu Chemical Co. (Tianjin, China).

Hexamethylenetetramine (HMTA), nitric acid (HNO3) and 1,4-benzoquinone (BQ) were

obtained from Yuanli Chemical Co. (Tianjin, China). L-glutamic acid, L-cysteine and

L-tryptophan were purchased from Sigma-Aldrich. Ciprofloxacin (CIP) was purchased from

Heowns Biochemical Technology Co. (Tianjin, China). Cephalexin (CEL), tert-butanol

(t-BuOH), triethanolamine (TEOA) and 3-aminopropyltriethoxysilane (APTES) were

purchased from Aladdin Industrial Corp. (Shanghai, China). All of the other chemicals, such

as acetonitrile and formic acid, were of chromatographic grade and were obtained from

commercial sources.

2.2. Synthesis of the N,S-CQDs

N,S-CQDs were synthesized via the hydrothermal treatment of L-cysteine with nitric acid

as a carbonization agent and ethylene glycol as a passivation agent. Briefly, 6.25 mL of HNO3

(68 wt%) was mixed with 4.75 mL of deionized water and 9.00 mL ethylene glycol. Then, 1.5

g of L-cysteine was dissolved in the as-prepared solution (20 mL). The obtained mixture was

stirred vigorously for 10 min and then transferred into a Teflon autoclave, heated to 180 C

for 10 h, and cooled to room temperature (25 C). The undispersed solids were removed by

centrifugation (10,000 rpm, 1 h). Then, the supernatant was dialyzed (MWCO 3500) to

remove any impurities in the CQDs sample. In contrast, the compared CQDs sample was
obtained through the same procedure except for the addition of ethylene glycol (denoted as

b-CQDs). The other two kinds of CQDs samples were prepared via L-glutamic acid and

L-tryptophan, which replaced the L-cysteine as carbon sources (denoted as Glu-CQDs and

Try-CQDs, respectively).

2.3. Synthesis of the ZnO/N,S-CQDs nanoflowers

Briefly, 0.669 g of Zn(NO3)2·6H2O was fully dissolved in 75 mL of deionized water, and

then 0.315 g of HMTA was dropped into the above solution and dissolved completely.

Another 0.9 g of NaOH was dissolved into 37.5 mL of deionized water, followed by dropwise

addition into the as-prepared mixture. Then, the N,S-CQDs after dialysis treatment was added

into the mixture under different mass ratios of ZnO. The resulting solution was vigorously

stirred at a speed of 550 rpm for 10 min at room temperature (25 C), then transferred into a

Teflon-sealed autoclave, and heated to 100 C for 12 h. The resulting products were

centrifuged, washed twice with deionized water and dried overnight. The different mass

ratios of N,S-CQDs to ZnO were prepared at 0.27, 0.41, 0.54 and 0.67 (denoted as

ZnO/N,S-CQDs0.27, ZnO/N,S-CQDs0.41, ZnO/N,S-CQDs0.54 and ZnO/N,S-CQDs0.67,

respectively). In contrast, the pure ZnO was prepared by the same procedure except for the

addition of N,S-CQDs. Moreover, three kinds of compared hybrid materials were also

synthesized via the addition of b-CQDs (denoted as ZnO/b-CQDs), Glu-CQDs (denoted as

ZnO/Glu-CQDs) and Try-CQDs (denoted as ZnO/Try-CQDs), respectively.

2.4. Evaluation of photocatalytic performance

To evaluate the photocatalytic properties of the as-prepared hybrid materials,

photodegradation experiments of multiple organic pollutants were performed under


irradiation of visible light (300 W xenon lamp with a 400 nm filter, 273.1 mW/cm2), NIR

light (300 W xenon lamp with a 700 nm filter, 109.6 mW/cm2) and simulated sunlight (300 W

xenon lamp with an AM1.5G filter, 264.5 mW/cm2). The typical reaction was as follows: 10

mg of the ZnO/N,S-CQDs sample was put into a photocatalytic device containing 25 mL of a

2×10-5 M organic pollutant solution. The resulting mixture was stirred for at least 30 min in

the dark to facilitate the adsorption equilibrium of the system. The distance from the reactants

to the lamp was maintained at 20 cm. To monitor the photodegradation efficiency, the

samples were collected from the as-prepared suspension, centrifuged to remove the solid

catalyst and then analyzed by using a UV-visible spectrophotometer at different time

intervals.

2.5. Application of the ZnO/N,S-CQDs nanoflowers in continuous flow reactor

The ZnO/N,S-CQDs nanoflowers were connected into the inner walls of the quartz

capillary tubes via the APTES coupling agent. At first, the quartz capillary tubes were

successively pretreated with 1 M HCl and 1 M NaOH at 40 C, and then 4.84 mL of APTES

was mixed with 145.16 mL of ethanol. The quartz capillary tubes were immersed in the

as-prepared solution with stirring for 3 h at 40 C. Then, the modified quartz capillary tubes

were stirred in the precursor solution of ZnO/N,S-CQDs (following the same procedure as

2.3). After the hydrothermal reaction at 100 C for 12 h, the quartz capillary tubes were

washed with deionized water many times and dried overnight. The quartz capillary tubes

were optimized by different inner diameters of each tube and flow rates of liquid (as shown in

Table S3).

2.6. Active species trapping experiments


To detect the active species generated during the photocatalytic reaction of the

ZnO/N,S-CQDs, various scavengers, including tert-butanol (t-BuOH), triethanolamine

(TEOA) and 1,4-benzoquinone (BQ), were introduced into the solution of CIP. The

-
hydroxide free radical ( ), holes (h+) and superoxide radical anion ( ) were trapped by

t-BuOH, TEOA and BQ, respectively.

2.7. Photodegradation of organic pollutants in actual water

Three kinds of actual water samples were respectively collected from the Tianjin

University Lake (TJU Lake), Tianta Lake and Haihe River (in Tianjin, China). Before the

degradation experiments, the water samples were filtered via microporous membranes (50

mm, 0.45 μm) to remove the suspended particulate matter. Multiple organic contaminants,

including MB, CIP and CEL, were dissolved into different actual water to form single, dual

or triple pollutant systems. Their photocatalytic performances were evaluated through the

same procedure as described in 2.4.

2.8. Characterization

Scanning electron microscopy (SEM) images were recorded using a field-emission

scanning electron microscope (FESEM, S4800, Hitachi High-Technologies Co., Japan) at an

acceleration voltage of 3 kV. Transmission electron microscopy (TEM) analyses were

performed using a field-emission transmission electron microscope (JEM2100F, JEOL,

Japan) at an accelerating voltage of 120 kV. Elemental analysis was conducted with an

energy dispersive X-ray spectrometer (EDS) equipped in the S-4800 FESEM at an

accelerating voltage of 10 kV. X-ray photoelectron spectroscopy (XPS) was performed with a

PHI1600 ESCA System (PerkinElmer, Waltham, MA). Fluorescence measurements were


carried out using a Cary Eclipse fluorescence spectrophotometer (Agilent Technologies, Inc.,

Santa Clara, CA). Diffuse reflectance spectra of the samples were recorded on a Perkin Elmer

Lambda 750 spectrophotometer by using BaSO4 as the 100% reflectance standard.

UV-visible spectra were recorded on a Pgeneral TU-1810 spectrophotometer (Purkinje Inc.,

Beijing, China) equipped with quartz cells. FT-IR spectra were collected on an FTIR

spectrophotometer (Nicolet 6700) using KBr disks at room temperature. The structural

characterization for the samples used an X-ray powder diffractometer (XRD, D/MAX-2500,

Japan) with Cu Kα radiation. The photoluminescence (PL) investigation was executed by

using a Fluorolog3 photoluminescence spectrometer (Horiba Jobin Yvon, America) with an

excitation wavelength of 325 nm. Nitrogen-adsorption experiments were performed at -196

°C on a Gemini VII 2390 apparatus. Brunauer-Emmett-Teller (BET) specific surface areas

were calculated based on the adsorption isotherms. The pore-size distribution was calculated

from the adsorption branch by using the BJH (Barrett-Joyner-Halenda) method. A TOC

analyzer (TOC-VCPH E200 V, Shimadzu Co., Japan) was employed to evaluate the

mineralization efficiency of the reaction solutions. The optical power density data were

recorded by an optical power meter (CEL-NP2000, Ceaulight Co., China). The zeta potential

measurement was performed on a NANO ZS ZEN3600 apparatus. The molecular structure of

N,S-CQDs was measured via the MALDI-TOF MS technique (Autoflex tof/tofIII, America).
3. Results and discussion

3.1. Structural characterization of the ZnO/N,S-CQDs nanoflowers

Scheme 1 shows the synthesis process of ZnO/N,S-CQDs nanoflowers. Specially, the

N,S-CQDs were embedded into the ZnO precursor solution and then grew via a one-pot

hydrothermal process. Due to the introduction of L-cysteine as a carbon source in the CQDs,

some -NH2 and -SH groups were exposed on their surfaces [23, 24]. Then, EG was employed

as a modifier to eliminate the superficial defects of CQDs. Theoretically, as an intermediate

bridge, the hydroxy groups in glycol units could connect with the carboxyl groups in

L-cysteine and simultaneously link to each other for the formation of polyethylene glycol or

oligo(ethylene glycol) [25] on the surfaces of N,S-CQDs. The molecular structure of

N,S-CQDs was characterized via MALDI-TOF MS technique, as shown in Fig. S1. The three

main protonation peaks, M/Z=1443, 1575, and 1751, in N,S-CQDs were probably from the

oligo(ethylene glycol) (the detailed results and discussion is shown in the supporting

information). Therefore, after the subsequent hydrothermal treatment, a novel

ZnO/N,S-CQDs heterostructured nanoflower embedded by the N,S-CQDs was formed.


Scheme 1. Schematic of the synthesis of ZnO/N,S-CQDs nanoflowers.

As shown in Fig. 1a, the ZnO/N,S-CQDs have an average diameter of 1.62 µm and show a

flower-like configuration composed of nanosheets with an average thickness of 38.67 nm.

Moreover, the sizes of the N,S-CQDs were 3-5 nm (Fig. 1c). The TEM images in Fig. 1d

show that the N,S-CQDs were embedded into the ZnO nanoflower according to the

interplanar spacings of 0.32 nm for CQDs and the surrounding spacings of 0.26 nm for ZnO

[26]. The distributions of elements were also investigated by the EDS spectrum of

ZnO/N,S-CQDs nanoflowers as shown in Fig. S3. In addition to zinc (Zn) and oxygen (O),

the carbon (C), sulfur (S) and nitrogen (N) were also distributed in the selected area of the

nanoflower. Due to the absence of them in the original ZnO, the S and N were mainly derived

from the N,S-CQDs embedded in the ZnO nanoflower.


Fig. 1. a-b) SEM images of the ZnO/N,S-CQDs0.54, TEM image of c) N,S-CQDs and d)

ZnO/N,S-CQDs0.54 (insets are the corresponding TEM images of ZnO/N,S-CQDs0.54).

Furthermore, a series of ZnO/N,S-CQDs nanoflowers with different mass ratios of

N,S-CQDs were synthesized, and the samples were denoted as ZnO/N,S-CQDsx (x=0.27,

0.41, 0.54 and 0.67). The light-harvesting capacities of the ZnO/N,S-CQD composites were

evaluated by the diffuse-reflectance spectra (Fig. 2a). It was found that the absorbance range

of ZnO was mainly in the ultraviolet regions below 400 nm, while the absorption edges of

ZnO/N,S-CQDs composites were integrally improved with increasing of N,S-CQDs. The

adsorption intensity of ZnO/N,S-CQDs0.54 reached the highest value among them, in which

the UV, visible and NIR light absorptions (λ>400 nm and λ>700 nm) were all involved. Fig.

2b shows the plots of (Ahν)1/2 versus photon energy (hν) via transformation of the

diffuse-reflectance spectra. The band-gap energy (Eg) (treated via the extrapolation method)
of ZnO/N,S-CQDs0.54 was observably decreased from 3.05 eV (of ZnO) to 2.70 eV.

Moreover, the FTIR spectra of N,S-CQDs, ZnO and ZnO/N,S-CQDs0.54 were recorded. As

shown in Fig. 2c, the bands at 3430 and 1037 cm-1 in the spectrum of N,S-CQDs respectively

represented the O-H and C-O stretching vibrations of glycol units or L-cysteine in the

N,S-CQDs [25]. The peaks at approximately 1205 cm-1 were attributed to the C-N and partly

C-S stretching vibrations from L-cysteine. Moreover, the peak at 1700 cm-1 corresponded to

the -C=O of carbonyl groups, and the bands at 2925 and 2964 cm-1 indicated the presence of

alkyl chain groups (e.g., -CH2- and -CH3) in the N,S-CQDs. As shown in Fig. 2d, the

intensities of the broad bands at 3425 and 1631 cm-1 of ZnO/N,S-CQDs0.54 were enhanced

when compared with that of ZnO. In particular, the peak at 3425 cm-1 was mainly attributed

to O-H stretching, while the band in 1631 cm-1 was possibly due to the N-H stretching

vibration peak in acid amides and -C=O groups [27, 28]. In addition, C-O vibrations in

N,S-CQDs were also observed in the spectrum of ZnO/N,S-CQDs0.54 at 1263 and 1072 cm-1.

The FTIR results provided further evidence in support of the embedding of N,S-CQDs in the

ZnO nanoflowers.
Fig. 2. a) Diffuse-reflectance spectra and b) plot of (Ahν)1/2 versus (hv) of ZnO and the

ZnO/N,S-CQDs with different mass ratios of N,S-CQDs, FTIR spectra of c) N,S-CQDs, d)

ZnO and ZnO/N,S-CQDs0.54.

The XPS analysis was further employed to evaluate the chemical composition of

ZnO/N,S-CQDs. It was found that multiple elements, including Zn2p, C1s, O1s, N1s and S2p

appeared on the full spectrum of ZnO/N,S-CQDs (Fig. 3). The specific element contents are

shown in Table S2, these results showed the presence of N1s and S2p in ZnO/N,S-CQDs0.54

but absence of signals in the ZnO, which further demonstrated the successful preparation of

N,S-CQDs in the ZnO/N,S-CQDs nanocomposites. In addition, the high-resolution spectrum


of C1s was fitted with four peaks at 284.1, 285.2, 286.6 and 288.1 eV and attributed to the

C-C (sp2), C-N, C-O and C=O bonds, respectively, mainly in glycol units and the carbonyl

structures of N,S-CQDs (Fig. 3c) [25]. As shown in Fig. 3d, three peaks can be decomposed

from the high resolution spectrum of O1s. The typical bond of -O-Zn-O- in ZnO material was

mainly at 530.0 eV, and the oxygen-containing groups of C-O and (C=O or O-H) in

ZnO/N,S-CQDs0.54 were at 531.0 and 532.4 eV, respectively [20, 26].

Fig. 3. a) XPS spectrum and the high resolution spectra of b) Zn 2p, c) C 1s and d) O 1s of

ZnO/N,S-CQDs0.54.

3.2. Photocatalytic activity of the ZnO/N,S-CQDs nanoflowers under visible and NIR

light irradiation

To evaluate the photocatalytic performance of the ZnO/N,S-CQDs nanoflowers, we


employed a variety of mimic organic pollutants containing MB, RhB, and MG for their

photodegradation under visible and NIR light. As shown in Fig. 4a, the degradation efficiency

of MB was 79% after 120 min under visible light by the ZnO/N,S-CQDs0.54, while only 9.3%

and 22.8% for N,S-CQDs and ZnO, respectively. It was worth mentioning that the

ZnO/N,S-CQDs0.54 showed a better photodegradation performance on MG compared with

other pollutants. Specifically, 85.4% of MG was degraded after 60 min under visible light

(Fig. 4a), and 72.8% of MG was decomposed after 180 min under NIR light (Fig. 4c).

Moreover, some other pollutants, including RhB and MB, were also partly degraded under

the irradiation of NIR light by the ZnO/N,S-CQDs0.54, which sharply contrasted with ZnO

and demonstrated the significance of the N,S-CQDs.

To gain insight into the photodegradation efficiency of organic dyes, we investigated the

adsorption of dyes on the ZnO/N,S-CQDs0.54 in the dark of 30 min. As shown in Fig. S6, the

adsorption efficiencies of ZnO/N,S-CQDs0.54 were 2.71, 2.28 and 12.19% for MB, RhB and

MG, respectively. As we know, the adsorption is the prerequisite for the degradation of dyes.

Therefore, the higher photodegradation efficiency of MG was probably attributed to the

larger adsorption compared to MB and RhB. In addition, the structural differences of organic

dyes likely contribute to their different photodegradation efficiencies. For example, the

differences in the unstable N+, carboxyl and the aniline groups [29] between RhB and MB

probably influenced on their photodegradation.

We compared some previous studies concerning the photodegradation of MG under visible

light. For instance, the CdSe quantum dots-modified TiO2 photocatalysts showed a

degradation efficiency of 97.6% after 360 min [30]. Moreover, 90% of MG was degraded by
the ridium-doped ZnO nanoparticles after 120 min [31]. In addition, the previous

N-ZnO/C-dots nanoflowers degraded 85% of MG after 160 min under visible light [20].

According to the above reports, most of the structures in previous photocatalysts were 1D

nanoparticles, which tended to assemble together during the reaction. While the novel

ZnO/N,S-CQDs in this work showed a special structure of nanosheets-composed nanoflower,

which effectively avoided the assembly, meanwhile, exposed most of highly reactive facets

on their surfaces [32]. Furthermore, after the subsequent introduction of N,S-CQDs, the

intrinsic absorption ranges of light of the novel nanocomposite were effectively extended and

the ZnO/N,S-CQDs showed better photocatalytic activities under visible and NIR light

irradiation.

Correspondingly, the photodegradation dynamics of the pollutants were also investigated.

The relationship between ln t 0) and time followed a pseudo-first-order kinetic equation:

ln t 0 - (1)

where 0 and t are the concentration of organic pollutants at the beginning and after a

period of time, respectively. k is a pseudo-first-order kinetic constant and t is the irradiation

time.

As a result, from Fig. 4b and Fig. 4d, the k values of ZnO and ZnO/N,S-CQDs0.54 were

respectively 0.0021 and 0.0122 min-1 for the degradation of MB under visible light. Moreover,

the k values of ZnO/N,S-CQDs0.54 were obviously improved to 0.0292 min-1 under visible

light and 0.0063 min-1 under NIR light irradiation for the degradation of MG.
Fig. 4. Photocatalytic activity and the pseudo-first-order kinetic curves of the blank control,

ZnO, N,S-CQDs and ZnO/N,S-CQDs0.54 for photodegradation of various organic pollutants

under a-b) visible (400-2500 nm) and c-b) NIR (700-2500 nm) light irradiation.

3.3. Photocatalytic activity of the ZnO/N,S-CQDs nanoflowers under simulated sunlight

irradiation

Herein, the ZnO/N,S-CQDs nanoflowers were further employed to evaluate the

photocatalytic degradation of refractory antibiotics (CIP, CEL) under simulated sunlight

irradiation. As shown in Fig. 5a, 92.9% of CIP was degraded within only 20 min by the

ZnO/N,S-CQDs0.54, and the degradation efficiency of CEL was 86.7% after 50 min. As

summarized in Table S1, the ZnO/N,S-CQDs showed a better photocatalytic efficiency than
the traditional photocatalysts and exhibited a comparative or even better performance than

some of the new-typed photocatalysts for degradation of CIP under simulated sunlight.

Generally, the degradation efficiency was relative with the nature of substrate and the

reaction conditions. Due to the different pKa values of CIP and CEL (specifically, the pKa

value of CIP was 6.09 for the carboxylic group and 8.74 for the nitrogen on the piperazinyl

ring; the pKa values of CEL were 3.6, 5.3 and 7.1 for different groups), we investigated the

effect of pH values on the photodegradation efficiency of CIP and CEL. As shown in Fig. S7,

both CIP and CEL had a maximum photodegradation efficiency at pH 7.2 and the degradation

efficiency of CIP was much higher than that of CEL, which was probably attributed to the

higher structural stability of CEL compared to CIP [33, 34].


Fig. 5. a-b) Photocatalytic activity and the pseudo-first-order kinetic constants of

ZnO/N,S-CQDs0.54 for the photodegradation of different organic pollutants, c-d)

Photocatalytic stability of the ZnO/N,S-CQDs0.54 for the degradation of CIP with five cycling

runs under simulated sunlight irradiation.

Moreover, the recycling ability of ZnO/N,S-CQDs0.54 was also investigated via five cycles

of photocatalytic reaction. As shown in Fig. 5c, the photodegradation efficiency of CIP was

85.7% after 5 cycles, which demonstrated a good recyclability of the hybrid nanoflower. The

result was probably attributed to the effectively embedded pattern of N,S-CQDs inside the

ZnO nanoflowers, which restricted the loss of N,S-CQDs during the reaction. The SEM

images of ZnO/N,S-CQDs0.54 after the photocatalytic cycles are shown in Fig. S8, it was
found that there were no obvious differences from the original ZnO/N,S-CQDs0.54. To

demonstrate the stability of N,S-CQDs in the composite, ZnO/N,S-CQDs was dispersed in

the deionized water and the supernatant was collected after centrifugation for fluorescence

analysis. As shown in Fig. S9, there was no obvious change in the fluorescence intensity of

the dispersion solution after three cycles. The fluorescence intensities of the supernatant were

all very low. The results indicated that the ZnO/N,S-CQDs composite possessed a high

stability and can be reused for multiple processes. Moreover, the FTIR, BET and XPS

characterizations of ZnO/N,S-CQDs0.54 after the photocatalytic reactions for five cycles were

performed and the results were shown in Fig. S10 and Fig. S11 (the detailed results and

discussions were shown in the supporting information). Overall, the results indicated that

there was no obvious change on the structure of ZnO/N,S-CQDs0.54 during the recycling

process and it possessed a high stability.

In addition to the above batch reactions, we often need to solve problems by means of

continuous flow reactions in industry. Therefore, the ZnO/N,S-CQDs nanoflowers were

further loaded on supporters so that the wastewater could be treated following the water

current at the same time. Owing to enough light illumination required by the photocatalysts,

we employed the quartz capillary tubes as supporters to connect with the ZnO/N,S-CQDs0.54

nanoflowers in their inner walls. As shown in Fig. 6a, the APTES was employed as a

modifier to decorate -NH2 groups in the inner walls of quartz capillary tubes [35].

Simultaneously, due to the exposure of most -OH groups on the surfaces of

ZnO/N,S-CQDs0.54, they were easily connected to the modified tubes via hydrogen bonds

[36]. The SEM images (Fig. S12c) show that the inner wall of a quartz capillary tube was
successfully modified with ZnO/N,S-CQDs0.54 nanoflowers. Furthermore, four pieces of

quartz capillary tubes with the same sizes were installed parallelly as a novel continuous flow

reactor for the photodegradation of CIP under the influence of a constant-flux pump in the

system (Fig. S12a-b).

Fig. 6. a) Schematic of the connection of ZnO/N,S-CQDs0.54 to the inner walls of the quartz

capillary tubes within the continuous flow reactor. b) Photodegradation efficiency of CIP in

the continuous flow reactors with various conditions under simulated sunlight irradiation.

The effects of the inner diameters of tubes and the flow rates of liquid on the photocatalytic

efficiencies were investigated. The parameters of the reaction conditions are summarized in

Table S3. As shown in Fig. 6b, the degradation efficiency of CIP was 42.1% after 90 min

when the flow rate of liquid was at 2 mL/min. While the flow rates increased to 5 and 8

mL/min, the degradation efficiency of CIP decreased to 34.2% and 27.6%, respectively,

which mainly attributed to the decreased retention time of liquid in the total tubes [36]. In

addition, the effect of inner diameter of tube was also investigated, as a result, when the inner

diameters exceeded 6 mm, the degradation activities of CIP were decreased obviously. It was

possibly attributed to the impaired contact of substrates with the photocatalysts at the inner

walls of quartz tubes and the weak receptivity of light on their surfaces [35, 36].
Therefore, the optimal flow rate of liquid and inner diameter of quartz capillary tubes were

respectively at 2 mL/min and 4 mm for the continuous flow reactor. Furthermore, the above

result was compared with that in a previous continuous flow device, in which the

mesostructured ZnO nanotubes were directly loaded on a glass microfiber filter via physical

absorption. It was found that only 12% of CIP was degraded after 120 min under simulated

sunlight [37]. In contrast, the continuous flow reactor composed by the novel ZnO/N,S-CQDs

photocatalyst in this work was more beneficial for the photodegradation of CIP.

3.4. Photocatalytic activity of the ZnO/N,S-CQDs nanoflowers under natural sunlight

irradiation

Moreover, we employed the natural sunlight as a direct light source for degradation of CIP

under the existence of ZnO/N,S-CQDs material. In this study, the photocatalytic reactions

were performed in two days with different light intensities of sunlight, in which the optical

power densities were severally measured as approximately 98.4 and 76.8 mW/cm2. As shown

in Fig. 7b, 85.8% of ciprofloxacin was degraded after 50 min by the ZnO/N,S-CQDs

nanoflowers under natural sunlight with an optical power density of 98.4 mW/cm2. The

degradation ratio decreased slightly to 79% when the optical power density was 76.8

mW/cm2. Recently, some photocatalysts were developed for the degradation of antibiotics in

water under natural sunlight irradiation. For example, Kumar et al. [38] reported the

quaternary magnetic BiOCl/g-C3N4/Cu2O/Fe3O4 heterojunction, which degraded 92.1% of

sulfamethoxazole after natural sunlight irradiation for 120 min. Wang et al. [39] designed a

ternary Ag3PO4/r-GO/g-C3N4 composite showing a degradation efficiency of 98% in the


treatment of norfloxacin for 30 min under intensive sunlight. In addition, the N-doped ZnS

microsphere was also reported for the removal of CIP from water, achieving 42% of

degradation efficiency after natural sunlight irradiation for 150 min [40]. In comparison with

these g-C3N4 based photocatalysts, the ZnO/N,S-CQDs exhibited a lower degradation

efficiency, e.g., 85.8% for CIP after 50 min of sunlight irradiation, however, which is higher

than that of N-doped ZnS microspheres.

Fig. 7. a) The photograph of photocatalytic device outdoors (in which 30 mg of

ZnO/N,S-CQDs0.54 and 50 mL of 2×10-5 M CIP solution were put into the system). b)

Photocatalytic activity of the ZnO/N,S-CQDs0.54 for the photodegradation of CIP under

natural sunlight with different intensities.

Furthermore, the reaction pathways and intermediates during the photodegradation of CIP

were investigated via HPLC-MS measurement. The possible molecular and structural

formulas were inferred and summarized in Table S4. In general, according to the previous

reports and corresponding inferences, two possible CIP degradation pathways were deduced

as shown in Fig. 8. For pathway I, the specific active site in the piperazine ring of CIP was
substituted by the -OH to form a structure of m/z = 348 under the main effect of hydroxide

free radicals ( ) [41, 42]. Furthermore, the piperazine ring group was opened and

underwent decarboxylation to form an intermediate with a five-membered ring structure with

m/z=274 [41]. Moreover, some intermediates with smaller molecular weights (containing

m/z=185 and 137) were detected in the system after subsequently advanced oxidation. In

addition, from pathway II, the -F group in CIP was substituted by -OH via attack of the

-
superoxide radical anions ( ) and the hydroxide free radicals ( ), resulting in the

generation of a defluorination product with m/z=330 [6, 42]. Afterwards, the decyclization of

the piperazine ring occurred and followed by a dealkylation, leading to the formation of the

intermediate with m/z value of 304 and its relevant derivative of 302. Moreover, the

subsequent products containing m/z=261, 244 and 228 were detected as a result of deeper

oxidation of deamination, dehydroxylation and deoxygenation. Finally, the intermediates

were promising for production of some smaller molecules, such as CO2, H2O and so on.

CO2
H2O
F-
NO3-

Fig. 8. Possible reaction pathways of CIP degraded by the ZnO/N,S-CQDs0.54 after natural

sunlight irradiation for 50 min (at an optical power density of 98.4 mW/cm2).
3.5. Photocatalytic mechanisms of the ZnO/N,S-CQDs nanoflowers

To construct a high-efficiency hybrid photocatalyst, we often need to evaluate some

internal characteristics of the dopants and investigate the photocatalytic mechanisms more

deeply [18]. Herein, we prepared the bare CQDs without EG modification and then

embedded them into the ZnO to get another hybrid material (which denoted as ZnO/b-CQDs)

for comparison. Specially, from the SEM images in Fig. S13, the ZnO/b-CQDs0.54 showed

some spindle nanorods instead of the nanosheets in ZnO/N,S-CQDs0.54. From Fig. 9b, the

photodegradation efficiency of CIP was only 8.6% after 20 min by the ZnO/b-CQDs0.54,

which was significantly lower than the value of 92.9% of ZnO/N,S-CQDs0.54 under simulated

sunlight irradiation.

Fig. 9. a) Schematic illustration of the construction of ZnO/b-CQDs and ZnO/N,S-CQDs. b)

The photodegradation efficiency of CIP by ZnO/b-CQDs0.54 and ZnO/N,S-CQDs0.54 under

simulated sunlight irradiation.

The structural mechanisms of ZnO/N,S-CQDs and ZnO/b-CQDs were compared with each

other in Fig. 9a. It is known that the direction of fastest intrinsic growth of ZnO was often the
[0001] crystal orientation. While under the influence of HMTA (due to its well-known

coordination properties), it acted as a kind of covering agent which linked in the (0001) plane

of ZnO nanorods during the process of growth [43]. Consequently, the ZnO nanosheets were

formed and with a majority of 0 0 facets. Subsequently, a kind of nanosheet-composed

nanoflower was obtained as a result of a tendency to minimum G (free Gibbs energy) of the

whole system [32]. In this work, when the b-CQDs were introduced into the system, the -OH

and -NH2 groups in the b-CQDs tended to preferentially link with the -OH on the surfaces of

ZnO via hydrogen bonds (Fig. 9a). Therefore, the function of HMTA was impacted in some

degree due to the connective competition of active sites between HMTA and the formation of

hydrogen bonds [28]. As a result, the superficial active sites on ZnO were occupied by the

hydrogen bonds partly that the HMTA could not adequately work at this time. Therefore, the

nanorod structures were formed in the ZnO/b-CQDs material. Correspondingly, when the

N,S-CQDs (with EG modification) were used, owing to the steric hindrance effects [25, 44],

only a few terminal -OH groups could link with ZnO for the formation of hydrogen bonds

that the competition of active sites was weakened a lot. Consequently, the

nanosheet-composed nanoflowers were formed in ZnO/N,S-CQDs material.

It was worth noting that the above explanations could also demonstrate the phenomenon in

Fig. S14. When the mass ratio of N,S-CQDs to ZnO increased to 0.67, the structure of

ZnO/N,S-CQDs gradually changed into the nanorods-constructed nanoflowers. The result

suggested that when the amount of N,S-CQDs increased to a certain degree, the connective

competition of active sites between HMTA and hydrogen bonds was still strong so that few

HMTA molecules were capped on the ZnO surfaces. As a result, the nanorods-composed
nanoflowers were formed [28]. Therefore, owing to the significance of highly reactive facets

for photocatalysts, the surface modification of CQDs with EG and the moderate amount of

N,S-CQDs were all significant for the construction of high-efficiency ZnO/CQDs hybrid

nanoflowers.

Fig. 10. a) Diffuse reflectance spectra, b) photoluminescence (PL), c) diagram of specific

surface areas, and d) photocatalytic activity (concerning photodegradation of CIP under

simulated sunlight irradiation) of different types of nanoflowers.

In addition, the other two EG-modified CQDs, Try-CQDs and Glu-CQDs, were prepared

using L-tryptophan and L-glutamic acid as the carbon sources of CQDs respectively.
Afterwards, the ZnO/Try-CQDs and ZnO/Glu-CQDs were synthesized following the same

procedures as ZnO/N,S-CQDs (the corresponding SEM images of the four types of

nanoflowers were shown in Fig. S18). As shown in Fig. 10a, ZnO/N,S-CQDs0.54 showed the

highest absorption intensity within the visible to NIR light ranges when compared with the

ZnO/Try-CQDs0.54 and ZnO/Glu-CQDs0.54. According to the PL spectra in Fig. 10b, the

minimal fluorescence intensity was presented on the ZnO/N,S-CQDs0.54, while the ZnO

showed the highest value. As we know, the fluorescence of a photocatalyst is often produced

by the recombination of photoelectrons and vacancies, therefore, the ZnO/N,S-CQDs with

lower fluorescence intensity showed a slower recombination rate of electrons and holes [32].

Moreover, the specific surface areas of the ZnO and ZnO/CQDs nanoflowers were recorded

in Fig. 10c, in which the ZnO/N,S-CQDs0.54 showed the highest specific surface areas among

the hybrid materials (Fig. S19 and Table S5). As expected, the photocatalytic activity of the

ZnO/CQDs nanoflowers were all superior to the primary ZnO (Fig. 10d). The degradation

efficiency of CIP under simulated sunlight showed an obvious trend of ZnO/N,S-CQDs0.54 >

ZnO/Glu-CQDs0.54 > ZnO/Try-CQDs0.54 > ZnO.

Furthermore, to investigate the photocatalytic mechanisms of the ZnO/N,S-CQDs

nanoflowers, the active species trapping experiments, the photocurrent responses of hybrid

nanoflowers, and the up-converted photoluminescence property of N,S-CQDs (Fig. S20)

were performed. In this study, different scavengers, including t-BuOH, TEOA and BQ were

added into the photocatalytic systems of CIP, in which the hydroxide free radicals ( ),

-
holes (h+) and superoxide radical anions ( ) were respectively trapped by t-BuOH, TEOA

and BQ. As shown in Fig. 11a, 38.8%, 12.4% and 1.6% of CIP were degraded by
ZnO/N,S-CQDs0.54 in the presence of t-BuOH, TEOA and BQ, respectively, which was much

lower than that (92.9%) in the absence of scavengers. It suggested that the three kinds of

-
active species all played certain roles in the photocatalytic reaction. In general, the was

evidently a main active specie during the process of photodegradation.

As shown in Fig. 11c, the photocurrent responses of the hybrid nanoflowers showed cyclic

alternations along with the on-off control of light [11, 19]. In particular, the highest transient

photocurrent presented on the ZnO/N,S-CQDs0.54 among the hybrid nanoflowers, indicating

that it possessed much higher transfer efficiency of photogenerated electrons and holes than

the others. In addition, the EIS data were also recorded in Fig. 11d, the arc radius on the EIS

Nyquist plot of ZnO/N,S-CQDs0.54 was smaller than that of ZnO/Glu-CQDs0.54,

ZnO/Try-CQDs0.54 and ZnO. The results indicated that the N,S-CQDs were more effective for

the acceleration of interfacial charge transfer efficiency, thus improving the photocatalytic

activity of the ZnO/N,S-CQDs nanoflowers [45, 46].


Fig. 11. a-b) Trapping experiments of active species during the photocatalytic degradation of

CIP by ZnO/N,S-CQDs0.54, c) Transient photocurrent responses and d) Electrochemical

impedance spectra (EIS) Nyquist plots of different types of nanoflowers under simulated

sunlight irradiation.

Overall, we summarized some kinds of hybrid photocatalysts composed of CQDs derived

from different carbon sources in Table S6. To date, various carbonaceous organic materials

have been employed for the construction of CQDs. However, when the CQDs were doped

into photocatalysts, some essential influencing factors should be taken into consideration.

The key points were as follows: 1) the photoinduced electron transfer and the excellent

internal characteristics of CQDs, especially for the up-converted luminescence property; 2)

the connection modes between basal photocatalysts and the CQDs, such as the doping,
embedding, chemical bonding and electrostatic absorption method and so on; 3) what is more,

the appropriate surface modification of CQDs for better combination with the basal

photocatalysts. To date, some reports involving the first two points have been investigated,

for instance, the CQDs derived from citric acid were added to Bi2WO6 nanosheets during the

hydrothermal process, which showed an improved photocatalytic activity ascribed to the

up-converted PL and electron transfer properties of the CQDs [19]. Additionally, Li et al.

constructed a novel CQDs/Cu2O composite by a one-step ultrasonic treatment method in

which the CQDs resulting from glucose showed efficient photocatalytic activity based on the

combination of light reflecting ability between Cu2O and the CQDs [47]. However, regarding

to the third point mentioned above, there have not been relevant studies. In this work, it was

found that the surface modification of N,S-CQDs was crucial for improvement of degradation

efficiency of the hybrid photocatalysts when compared to the b-CQDs.

Herein, the N,S-CQDs not only covered the good advantages in previous reports but also

connected more efficiently with the basal ZnO. As shown in Scheme 1, after the addition of

N,S-CQDs during the hydrothermal reaction, a kind of nanosheets-composed nanoflower was

formed. Due to the good electron storage capacities that carbon nanomaterials often

possessed, the N,S-CQDs acted as an important electron mediator that transferred the

electrons from the ZnO and effectively postponed the recombination of photo-generated

electrons and holes in the hybrid material [48, 49]. What is more, owing to the good

up-converted luminescence property, the N,S-CQDs could emit the low-wavelength visible

light when they absorbed the NIR light. Therefore, the intrinsic absorption ranges of light for

the hybrid ZnO/N,S-CQDs material were extended obviously that it could degrade organic
pollutants not only under the visible light but also the NIR light irradiation.

3.6. Photodegradation of organic pollutants in actual water

Finally, we detected the photodegradation efficiency of ZnO/N,S-CQDs material in actual

water. Specifically, three kinds of water samples were respectively collected from Tianjin

University Lake (TJU Lake), Tianta Lake and the Haihe River. Then CIP, CEL and MB were

severally added to the water samples to mimic different types of wastewater containing single,

dual and triple pollutants. Afterwards, the pollutants were degraded by the ZnO/N,S-CQDs0.54

nanoflowers under simulated sunlight irradiation. As shown in Fig. 12a, the three wastewater

samples containing CIP had similar kinetic constants (~0.02 min-1), which is lower than that

in pure water (~0.13 min-1, Fig. 5b), while MB in the Haihe River had a higher k value

compared with CIP and CEL. When both CIP and MB were included in wastewater, 73.6% of

CIP and 95.1% of MB were degraded under simulated sunlight irradiation for 80 min and 40

min, respectively (Fig. 12b). For the mixture of CIP and CEL in wastewater, 71.7% of CIP

and 70% of CEL were degraded after 120 min (Fig. 12c). As expected, in the CIP-CEL-MB

triple pollutants system, the photocatalytic efficiency of the antibiotics still remained above

60% after 120 min, and 94.9% of the initial MB was degraded within 40 min (Fig. 12d). In

general, the ZnO/N,S-CQDs exhibited good photocatalytic performance for simultaneous

degradation of antibiotics and organic dyes in actual water.


Fig. 12. Photodegradation of various organic pollutants by ZnO/N,S-CQDs0.54 in the three

types of actual water via a) single, b-c) dual and d) triple pollutant systems.

4. Conclusions
In summary, we have successfully synthesized a new 3D ZnO/N,S-CQDs nanoflower via

one-pot hydrothermal process, in which the N,S-CQDs was prepared via the hydrothermal

treatment using L-cysteine as a carbon source and ethylene glycol as a passivation agent. The

surface functionalization, electron transfer, up-converted luminescence properties of

N,S-CQDs, together with the highly reactive facets of ZnO nanoflowers, made great

contributions to the enhanced photocatalytic activity of ZnO/N,S-CQDs under visible and

NIR lights. The as-prepared ZnO/N,S-CQDs effectively degraded antibiotics in water under

sunlight irradiation. In addition, the ZnO/N,S-CQDs exhibited high stability and good
recyclability, probably due to the embedment of N,S-CQDs into the ZnO nanoflowers.

Moreover, it is demonstrated that the ZnO/N,S-CQDs composite can degraded antibiotics and

organic dyes in actual water, achieving good photocatalytic performance under sunlight

irradiation. In view of the relatively low cost and facile one-pot synthesis, we believe that the

ZnO/N,S-CQDs nanoflowers have potential in the large-scale photocatalytic degradation of

environmental pollutants.

Acknowledgments
This work was supported by the National Natural Science Foundation of China (Nos.

21621004, 51773149 and 21777112) and Tianjin Municipal Science and Technology Bureau,

China (18YFHBZC00010).

Appendix A. Supplementary data


Supplementary data associated with this article can be found in the online version at

http://dx.doi.org/...
References
[1] R. Zhang, J. Tang, J. Li, Z. Cheng, C. Chaemfa, D. Liu, Q. Zheng, M. Song, C. Luo,

G. Zhang, Occurrence and risks of antibiotics in the coastal aquatic environment of

the Yellow Sea, North China, Sci. Total. Environ. 450-451 (2013) 197-204.

[2] S. Manzetti, R. Ghisi, The environmental release and fate of antibiotics, Mar. Pollut.

Bull. 79 (2014) 7-15.

[3] M. Qiao, G.G. Ying, A.C. Singer, Y.G. Zhu, Review of antibiotic resistance in China

and its environment, Environ. Int. 110 (2018) 160-172.

[4] I. Michael, L. Rizzo, C.S. McArdell, C.M. Manaia, C. Merlin, T. Schwartz, C. Dagot,

D. Fatta-Kassinos, Urban wastewater treatment plants as hotspots for the release of

antibiotics in the environment: a review, Water. Res. 47 (2013) 957-995.

[5] K.M. Lee, C.W. Lai, K.S. Ngai, J.C. Juan, Recent developments of zinc oxide based

photocatalyst in water treatment technology: A review, Water. Res. 88 (2016)

428-448.

[6] F. Wang, Y. Feng, P. Chen, Y. Wang, Y. Su, Q. Zhang, Y. Zeng, Z. Xie, H. Liu, Y. Liu,

W. Lv, G. Liu, Photocatalytic degradation of fluoroquinolone antibiotics using

ordered mesoporous g-C3N4 under simulated sunlight irradiation: Kinetics,

mechanism, and antibacterial activity elimination, Appl. Catal. B Environ. 227 (2018)

114-122.

[7] J.C. Durán-Álvarez, E. Avella, R.M. Ramírez-Zamora, R. Zanella, Photocatalytic

degradation of ciprofloxacin using mono-(Au, Ag and Cu) and bi-(Au-Ag and Au-Cu)
metallic nanoparticles supported on TiO2 under UV-C and simulated sunlight, Catal.

Today 266 (2016) 175-187.

[8] Y. Gan, Y. Wei, J. Xiong, G. Cheng, Impact of post-processing modes of precursor on

adsorption and photocatalytic capability of mesoporous TiO2 nanocrystallite

aggregates towards ciprofloxacin removal, Chem. Eng. J. 349 (2018) 1-16.

[9] H.H. Lin, A.Y. Lin, Photocatalytic oxidation of 5-fluorouracil and cyclophosphamide

via UV/TiO2 in an aqueous environment, Water. Res. 48 (2014) 559-568.

[10] J. Choi, S. Chan, H. Joo, H. Yang, F.K. Ko, Three-dimensional (3D) palladium-zinc

oxide nanowire nanofiber as photo-catalyst for water treatment, Water. Res. 101 (2016)

362-369.

[11] N. Li, J. Zhang, Y. Tian, J. Zhao, J. Zhang, W. Zuo, Precisely controlled fabrication of

magnetic 3D γ-Fe2O3@ZnO core-shell photocatalyst with enhanced activity:

Ciprofloxacin degradation and mechanism insight, Chem. Eng. J. 308 (2017)

377-385.

[12] Y. Hong, C. Li, G. Zhang, Y. Meng, B. Yin, Y. Zhao, W. Shi, Efficient and stable

Nb2O5 modified g-C3N4 photocatalyst for removal of antibiotic pollutant, Chem. Eng.

J. 299 (2016) 74-84.

[13] D. Mao, A. Yu, S. Ding, F. Wang, S. Yang, C. Sun, H. He, Y. Liu, K. Yu, One-pot

synthesis of BiOCl half-shells using microemulsion droplets as templates with highly

photocatalytic performance for the degradation of ciprofloxacin, Appl. Surf. Sci. 389

(2016) 742-750.

[14] Y. Hao, X. Dong, X. Wang, S. Zhai, H. Ma, X. Zhang, Controllable electrostatic


self-assembly of sub-3 nm graphene quantum dots incorporated into mesoporous

Bi2MoO6 frameworks: efficient physical and chemical simultaneous co-catalysis for

photocatalytic oxidation, J. Mater. Chem. A 4 (2016) 8298-8307.

[15] A.Y. Zhang, W.K. Wang, D.N. Pei, H.Q. Yu, Degradation of refractory pollutants

under solar light irradiation by a robust and self-protected ZnO/CdS/TiO2 hybrid

photocatalyst, Water. Res. 92 (2016) 78-86.

[16] H. Yuan, Y. Zhao, Y. Wang, J. Duan, B. He, Q. Tang, Sonochemistry-assisted

black/red phosphorus hybrid quantum dots for dye-sensitized solar cells, J. Power

Sources 410-411 (2019) 53-58.

[17] Y.C. Nie, F. Yu, L.C. Wang, Q.J. Xing, X. Liu, Y. Pei, J.P. Zou, W.L. Dai, Y. Li, S.L.

Suib, Photocatalytic degradation of organic pollutants coupled with simultaneous

photocatalytic H2 evolution over graphene quantum dots/Mn-N-TiO2/g-C3N4

composite catalysts: Performance and mechanism, Appl. Catal. B Environ. 227 (2018)

312-321.

[18] N. Shao, Z. Hou, H. Zhu, J. Wang, C.P. François-Xavier, Novel 3D core-shell

structured CQDs/Ag3PO4@Benzoxazine tetrapods for enhancement of visible-light

photocatalytic activity and anti-photocorrosion, Appl. Catal. B Environ. 232 (2018)

574-586.

[19] J. Wang, L. Tang, G. Zeng, Y. Deng, H. Dong, Y. Liu, L. Wang, B. Peng, C. Zhang, F.

Chen, 0D/2D interface engineering of carbon quantum dots modified Bi2WO6

ultrathin nanosheets with enhanced photoactivity for full spectrum light utilization

and mechanism insight, Appl. Catal. B Environ. 222 (2018) 115-123.


[20] S. Sharma, S.K. Mehta, S.K. Kansal, N doped ZnO/C-dots nanoflowers as visible

light driven photocatalyst for the degradation of malachite green dye in aqueous phase,

J. Alloy. Compd. 699 (2017) 323-333.

[21] J. Tian, Y. Leng, Z. Zhao, Y. Xia, Y. Sang, P. Hao, J. Zhan, M. Li, H. Liu, Carbon

quantum dots/hydrogenated TiO2 nanobelt heterostructures and their broad spectrum

photocatalytic properties under UV, visible, and near-infrared irradiation, Nano

Energy 11 (2015) 419-427.

[22] F. Wang, Y. Wang, Y. Feng, Y. Zeng, Z. Xie, Q. Zhang, Y. Su, P. Chen, Y. Liu, K. Yao,

W. Lv, G. Liu, Novel ternary photocatalyst of single atom-dispersed silver and carbon

quantum dots co-loaded with ultrathin g-C3N4 for broad spectrum photocatalytic

degradation of naproxen, Appl. Catal. B Environ. 221 (2018) 510-520.

[23] M. Li, M. Wang, L. Zhu, Y. Li, Z. Yan, Z. Shen, X. Cao, Facile microwave assisted

synthesis of N-rich carbon quantum dots/dual-phase TiO2 heterostructured

nanocomposites with high activity in CO2 photoreduction, Appl. Catal. B Environ.

231 (2018) 269-276.

[24] F. Li, Y. Li, X. Yang, X. Han, Y. Jiao, T. Wei, D. Yang, H. Xu, G. Nie, Highly

fluorescent chiral N-S-doped carbon dots from cysteine: Affecting cellular energy

metabolism, Angew. Chem. Int. Ed. Engl. 57 (2018) 2377-2382.

[25] Z. Wang, Y. Qu, X. Gao, C. Mu, J. Bai, Q. Pu, Facile preparation of oligo(ethylene

glycol)-capped fluorescent carbon dots from glutamic acid for plant cell imaging,

Mater. Lett. 129 (2014) 122-125.

[26] Y. Li, B.P. Zhang, J.X. Zhao, Z.H. Ge, X.K. Zhao, L. Zou, ZnO/carbon quantum dots
heterostructure with enhanced photocatalytic properties, Appl. Surf. Sci. 279 (2013)

367-373.

[27] J. Li, K. Liu, J. Xue, G. Xue, X. Sheng, H. Wang, P. Huo, Y. Yan, CQDs preluded

carbon-incorporated 3D burger-like hybrid ZnO enhanced visible-light-driven

photocatalytic activity and mechanism implication, J. Catal. 369 (2019) 450-461.

[28] H. Liu, B. Huang, Z. Wang, X. Qin, X. Zhang, J. Wei, Y. Dai, P. Wang, M.H.

Whangbo, Amino acid-assisted synthesis of ZnO twin-prisms and functional group's

influence on their morphologies, J. Alloy. Compd. 507 (2010) 326-330.

[29] T.S. Natarajan, M. Thomas, K. Natarajan, H.C. Bajaj, R.J. Tayade, Study on

UV-LED/TiO2 process for degradation of Rhodamine B dye, Chem. Eng. J. 169 (2011)

126-134.

[30] P. Wang, D. Li, J. Chen, X. Zhang, J. Xian, X. Yang, X. Zheng, X. Li, Y. Shao, A

novel and green method to synthesize CdSe quantum dots-modified TiO2 and its

enhanced visible light photocatalytic activity, Appl. Catal. B Environ. 160-161 (2014)

217-226.

[31] N. Babajani, S. Jamshidi, Investigation of photocatalytic malachite green degradation

by iridium doped zinc oxide nanoparticles: Application of response surface

methodology, J. Alloy. Compd. 782 (2019) 533-544.

[32] R. Shi, P. Yang, X. Song, J. Wang, Q. Che, A. Zhang, ZnO flower: Self-assembly

growth from nanosheets with exposed 00 facet, white emission, and enhanced

photocatalysis, Appl. Surf. Sci. 366 (2016) 506-513.

[33] A. Salma, S. Thoroe-Boveleth, T.C. Schmidt, J. Tuerk, Dependence of transformation


product formation on pH during photolytic and photocatalytic degradation of

ciprofloxacin, J. Hazard. Mater. 313 (2016) 49-59.

[34] N. Ajoudanian, A. Nezamzadeh-Ejhieh, Enhanced photocatalytic activity of nickel

oxide supported on clinoptilolite nanoparticles for the photodegradation of aqueous

cephalexin, Mat. Sci. Semicon. Proc. 36 (2015) 162-169.

[35] J. Li, F. Wu, L. Lin, Y. Guo, H. Liu, X. Zhang, Flow fabrication of a highly efficient

Pd/UiO-66-NH2 film capillary microreactor for 4-nitrophenol reduction, Chem. Eng. J.

333 (2018) 146-152.

[36] W. Gao, X. Zhang, X. Su, F. Wang, Z. Liu, B. Liu, J. Zhan, H. Liu, Y. Sang,

Construction of bimetallic Pd-Ag enhanced AgBr/TiO2 hierarchical nanostructured

photocatalytic hybrid capillary tubes and devices for continuous photocatalytic

degradation of VOCs, Chem. Eng. J. 346 (2018) 77-84.

[37] C. Bojer, J. Schöbel, T. Martin, M. Ertl, H. Schmalz, J. Breu, Clinical wastewater

treatment: Photochemical removal of an anionic antibiotic (ciprofloxacin) by

mesostructured high aspect ratio ZnO nanotubes, Appl. Catal. B Environ. 204 (2017)

561-565.

[38] A. Kumar, A. Kumar, G. Sharma, A.H. Al-Muhtaseb, M. Naushad, A.A. Ghfar, F.J.

Stadler, Quaternary magnetic BiOCl/g-C3N4/Cu2O/Fe3O4 nano-junction for visible

light and solar powered degradation of sulfamethoxazole from aqueous environment,

Chem. Eng. J. 334 (2018) 462-478.

[39] J. Wang, H. Chen, L. Tang, G. Zeng, Y. Liu, M. Yan, Y. Deng, H. Feng, J. Yu, L. Wang,

Antibiotic removal from water: A highly efficient silver phosphate-based Z-scheme


photocatalytic system under natural solar light, Sci. Total. Environ. 639 (2018)

1462-1470.

[40] L. Tie, R. Sun, H. Jiang, Y. Liu, Y. Xia, Y.Y. Li, H. Chen, C. Yu, S. Dong, J. Sun, J.

Sun, Facile fabrication of N-doped ZnS nanomaterials for efficient photocatalytic

performance of organic pollutant removal and H2 production, J. Alloy. Compd. 807

(2019) 151670.

[41] X. Zheng, S. Xu, Y. Wang, X. Sun, Y. Gao, B. Gao, Enhanced degradation of

ciprofloxacin by graphitized mesoporous carbon (GMC)-TiO2 nanocomposite: Strong

synergy of adsorption-photocatalysis and antibiotics degradation mechanism, J.

Colloid. Interface. Sci. 527 (2018) 202-213.

[42] X. Ao, W. Liu, W. Sun, M. Cai, Z. Ye, C. Yang, Z. Lu, C. Li, Medium pressure

UV-activated peroxymonosulfate for ciprofloxacin degradation: Kinetics, mechanism,

and genotoxicity, Chem. Eng. J. 345 (2018) 87-97.

[43] Y. Miao, H. Zhang, S. Yuan, Z. Jiao, X. Zhu, Preparation of flower-like ZnO

architectures assembled with nanosheets for enhanced photocatalytic activity, J.

Colloid. Interface. Sci. 462 (2016) 9-18.

[44] S. Liao, X. Zhao, F. Zhu, M. Chen, Z. Wu, X. Song, H. Yang, X. Chen, Novel S,

N-doped carbon quantum dot-based "off-on" fluorescent sensor for silver ion and

cysteine, Talanta 180 (2018) 300-308.

[45] J. Yu, D. Sun, T. Wang, F. Li, Fabrication of Ag@AgCl/ZnO submicron wire film

catalyst on glass substrate with excellent visible light photocatalytic activity and

reusability, Chem. Eng. J. 334 (2018) 225-236.


[46] K.S. Ranjith, R.B. Castillo, M. Sillanpaa, R.T. Rajendra Kumar, Effective shell wall

thickness of vertically aligned ZnO-ZnS core-shell nanorod arrays on visible

photocatalytic and photo sensing properties, Appl. Catal. B Environ. 237 (2018)

128-139.

[47] H. Li, R. Liu, Y. Liu, H. Huang, H. Yu, H. Ming, S. Lian, S.T. Lee, Z. Kang, Carbon

quantum dots/Cu2O composites with protruding nanostructures and their highly

efficient (near) infrared photocatalytic behavior, J. Mater. Chem. 22 (2012) 17470.

[48] H. Wang, Z. Wei, H. Matsui, S. Zhou, Fe3O4/carbon quantum dots hybrid nanoflowers

for highly active and recyclable visible-light driven photocatalyst, J. Mater. Chem. A 2

(2014) 15740-15745.

[49] M. Pirsaheb, A. Asadi, M. Sillanpää, N. Farhadian, Application of carbon quantum

dots to increase the activity of conventional photocatalysts: A systematic review, J.

Mol. Liq. 271 (2018) 857-871.


Graphical abstract
Highlights
 A new N,S-CQDs was synthesized using ethylene glycol as a passivation agent.

 A new 3D ZnO/N,S-CQDs nanoflower was synthesized via one-pot hydrothermal process.

 ZnO/N,S-CQDs possessed enhanced catalytic efficiency under visible and NIR lights.

 ZnO/N,S-CQDs effectively degraded antibiotics in water under sunlight irradiation.

You might also like