You are on page 1of 27

Journal Pre-proof

Dispersive solid-phase extraction of bisphenols migrated from plastic


food packaging materials with cetyltrimethylammonium
bromide-intercalated zinc oxide

Lili Lian , Xinhao Jiang , Junjiao Guan , Zhifu Qiu , Xiyue Wang ,
Dawei Lou

PII: S0021-9673(19)31088-X
DOI: https://doi.org/10.1016/j.chroma.2019.460666
Reference: CHROMA 460666

To appear in: Journal of Chromatography A

Received date: 8 July 2019


Revised date: 22 October 2019
Accepted date: 29 October 2019

Please cite this article as: Lili Lian , Xinhao Jiang , Junjiao Guan , Zhifu Qiu , Xiyue Wang ,
Dawei Lou , Dispersive solid-phase extraction of bisphenols migrated from plastic food packaging
materials with cetyltrimethylammonium bromide-intercalated zinc oxide, Journal of Chromatography A
(2019), doi: https://doi.org/10.1016/j.chroma.2019.460666

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.


Highlights

 ZnO@CTAB had been prepared by a one pot hydrothermal method.

 ZnO@CTAB was used as adsorbent for dispersive solid-phase extraction of BPs.

 The migration of BPA and BPAF from plastic packaging materials was test.

1
Dispersive solid-phase extraction of bisphenols migrated from

plastic food packaging materials with

cetyltrimethylammonium bromide-intercalated zinc oxide

Lili Lian, Xinhao Jia ng, Junjiao Guan, Zhifu Qiu, Xiyue Wang, Dawei Lou*

Department of Analytical Chemistry, Jilin Institute of Chemical Technology, No. 45

Chengde Street, Jilin 132022, PR China

Abstract

*
Corresponding author. Tel.: +86 432 63083163; fax: +86 432 63083163.
Email address: dwlou@hotmail.com (D.W. Lou).

2
The migration of bisphenols (BPs) can take place from plastic packaging materials into

freshly cooked takeaway food, especially at high temperatures. In this study,

cetyltrimethylammonium bromide-intercalated zinc oxide (ZnO@CTAB) was developed

and used to extract the migration of BPs (bisphenol A and bisphenol AF) from disposable

plastic materials to contained food simulates. Several experimental parameters that

influence extraction efficiency were investigated. Under optimal conditions, a sensitive

dispersive solid-phase extraction method based on ZnO@CTAB was proposed for the

analysis of BPs coupled with ultra-high performance liquid chromatography and

time-of-flight mass spectrometry. The method exhibited good linearity of calibration,

high recovery (97.63–109.33%), low limits of detection 0.027–0.030 μg L-1, and

acceptable precisions. The developed method was used to carry out a migration test from

two disposable plastic bags and a disposable plastic container, using distilled water at

100°C as a hot liquid food simulant. The migration concentrations of bisphenol AF was

found to be 0.42 μg L-1 and 0.86 μg L-1 for the two types of disposable plastic bags, and

the concentration of bisphenol A was 0.49 μg L-1 for disposable plastic container. The

proposed method was also applied to investigate the migration of BPAF from the

disposable plastic bags to different food simulants, revealing that the release of BPAF

levels depended on the polarity of the liquid food components.

Keywords: Dispersive solid-phase extraction; bisphenols; migration; plastic packaging

materials

1. Introduction

3
The consumption of takeaway food has been grown rapidly in recent years as a

consequence of modern lifestyles [1]. Takeaway food is often packaged with plastic

materials due to their convenience and competitive pricing. However, it has been reported

that polymerization reactions may not be completed during the manufacture of

polycarbonate plastic, and that some plastics components could be released to contacted

food. Bisphenol A (BPA), along with its similar derivative bisphenol AF (BPAF), are

extensively used as monomers in the manufacture of plastic packaging materials [2].

These bisphenol compounds (BPs) are estrogenic endocrine-disrupters that negatively

affect both humans and wildlife [3, 4]. BPA and BPAF have been frequently detected in

beverages, foodstuffs, and water that has been packaged in polycarbonate plastic

materials. Furthermore, studies of the migration of components from plastic materials

have shown that BP migration is seriously affected by the temperature of food [5]. To

avoid potential food risks to the consumer, it is extremely important to estimate the levels

of BPA and BPAF that migrate from polycarbonate plastic packaging materials into

freshly cooked takeaway food.

The primary techniques for BP analysis include the use of electrochemical sensors,

optical sensors, gas chromatography (GC) and high-performance liquid chromatography

(HPLC) coupled with various detectors. Of these techniques, HPLC coupled with mass

spectrometry (MS) has attracted much attention due to its effective separation and

accurate determination of various BP derivatives. Additional sample pretreatment steps

are needed before HPLC-MS analysis to eliminate or decrease matrix interferences and

improve the sensitivity of the method [6]. Dispersive solid-phase extraction (DSPE), a

new type of SPE technique, has also attracted much attention [7-9]. In DSPE, an

4
adsorbent is fully contacted with a target solution by vortex mixing and then separated by

centrifugation or passing through a filter membrane. After desorption with a small

volume of solvent, the recovered analytes are analyzed. DSPE avoids the high back

pressure, poor repeatability, and excessive solvent consumption of traditional SPE [10,

11]. In the DSPE procedure, the choice of adsorbent material plays a key role.

An abundant, low-cost and eco-friendly metal oxide, zinc oxide (ZnO) has attracted

much attention in adsorption and separation. ZnO exhibits good adsorption capability for

water soluble dyes [12] and antibiotics [13], due to its large surface area and abundant

active sites. However, the presence of negative charges and large numbers of -OH groups

on the surface results in ZnO exhibiting strong hydrophilicity, and, therefore, it is not an

efficient DSPE adsorbent for hydrophobic BPs. To further improve the adsorption

capacity of raw materials, their modification with other functional groups and materials is

envisaged. Cetyltrimethylammonium bromide (CTAB), a commonly used cationic

surfactant, has been adopted in the modification of bentonite [14], polystyrene [15] and

metal-organic frameworks (MOFs) [16]. In particular, it has been reported that CTAB

can change surface properties of nanocomposites from hydrophilic to hydrophobic [17].

In addition, CTAB has a positively charged oleophilic quaternary ammonium moiety that

can interact with the benzene rings of organic macromolecules through cation-π electron

interactions [18]. Herein, CTAB was selected to modify ZnO to improve its extraction

performance for hydrophobic BPs.

In the present study, CTAB-intercalated ZnO (ZnO@CTAB), was prepared by a

one-pot hydrothermal method. A sensitive analytical method for the detection of BPA

and BPAF was proposed in combination with ultra-high performance liquid

5
chromatography and time-of-flight mass spectrometry (UPLC/TOF-MS). Compared with

other commonly used DSPE adsorbents, ZnO@CTAB is inexpensive, simple to prepare,

and environmentally friendly. Moreover, the BP migration from disposable plastic bags

and disposable plastic container to hot liquid food simulants were assessed by the

developed method.

2. Experimental Methods

2.1. Materials and reagents

HPLC grade BPs (BPA and BPAF) and analytical grade zinc acetate [Zn(Ac)2•2H2O],

sodium citrate, urea, and CTAB were obtained from Aladdin Chemistry Co. Ltd.

(Shanghai, China). HPLC grade methanol (MeOH) and acetonitrile (MeCN) were

purchased from Fisher (Pittsburgh, PA, USA).

2.2. UPLC and Q-TOF-MS condition

BPA and BPAF were separated on an Acquity ultra-high performance liquid

chromatography (UPLC) system with a BEH C18 column (Waters, 50 mm × 2.1 mm

inner diameter, 1.7 μm) at 40°C. The injection volume was 10 μL. Chromatographic

separation and mass spectrometry conditions were illustrated in Supplementary data.

2.3. Characterization of ZnO@CTAB

The morphology of ZnO@CTAB was analyzed by scanning electron microscope

(SEM, JSM-7500F, JEOL, Japan) and transmission electron microscopy (TEM, JEOL

JEM-2100F). The functional groups of the prepared materials were certified by recording

6
the IR spectra in the range 4000–400 cm-1 using a Fourier transform infrared

spectrophotometer (FT-IR) (Vertex 80V, Bruker, Germany). The crystalline phase of

ZnO@CTAB was recorded on an X-ray diffractometer (XRD, D8 Advanced, Bruker,

Germany) equipped with a Cu-Kα radiation in the 2θ range of 10–80o. The nitrogen

adsorption-desorption isotherm was analyzed with a surface area analyzer (ASAP2000,

Micromeritics, USA). Thermal properties were measured by a thermal gravimetric

analyzer (TGA, Q500, Schimidt, Germany). Surface charges were tested by measuring

ζ-potential with a Nano Zetasizer (ZS/ZEN3690, Malvern, England).

2.4. Preparation of adsorbents

ZnO@CTAB was synthesized by a one pot hydrothermal method. Briefly, zinc

acetate (3 mmol), urea (6 mmol), sodium citrate (0.3 mmol) and CTAB (0.1–0.9 g) were

dissolved in water (75 mL). After maintaining sonication for 10 min, the mixture was

transferred into a 100 mL teflon-lined stainless steel reactor and heated at 120°C for 6 h.

The obtained white sample was washed with ultrapure water and ethanol several times,

and dried at 80°C under vacuum for 6 h to obtain ZnO@CTAB. For comparison,

mesoporous ZnO was synthesized by a similar hydrothermal method without the addition

of CTAB.

2.5. Adsorption property of ZnO@CTAB

Static adsorption experiments were conducted as follows: 5 mg of adsorbent was

added into 10.0 mL of BPs solution at various concentrations (20–80 mg L-1) and the

mixture incubated at room temperature for 2 h. Subsequently, the adsorbent was

7
separated by centrifugation, and the concentrations of BPs in the supernatant were

measured by UPLC. Langmuir and Freundlich isotherm models were used for evaluation

of the adsorption capacity of ZnO@CTAB. In addition, dynamic adsorption experiments

were also carried out at a BP concentration of 20 mg L-1.

2.6. DSPE procedure

The application of ZnO@CTAB as an adsorbent for the extraction of BPs was

conducted as follows. 10 mg of ZnO@CTAB was dispersed into a 20 mL glass vial

containing 10.0 mL of spiked analyte solution. After shaking at room temperature for 20

min, the analyte-loaded adsorbent was centrifuged for 10 min at 10,000 rpm. A mixture

of MeOH and MeCN (1:1 v/v, 2.5 mL) was used as the eluent to desorb the analytes. The

collected eluents were evaporated to dryness under a gentle nitrogen stream. Finally, the

analytes were re-dissolved in a mixture of MeOH and water (2:3 v/v, 1.0 mL) and

analyzed by UPLC/Q-TOF-MS.

2.7. Determination of migrated BPs from disposable plastic packaging

One purpose of this study was to evaluate the migration of BPs from disposable plastic

packaging to freshly cooked food, using ZnO@CTAB as an adsorbent. In China,

disposable plastic bags and disposable plastic containers are often used to package

takeaway food, such as noodles, soup, rice noodles, and spicy hot pots. Two types of

disposable plastic bags (denoted by disposal plastic bag-Ⅰ and disposal plastic bag-Ⅱ)

and a disposable plastic container were purchased from the local supermarket in Jilin

(China). Migration experiments from the disposable plastic packaging materials were first

8
performed using distilled water as food simulant at a constant temperature (100°C). 100.0

mL of hot water was transferred into the plastic packaging materials and maintained at

this temperature for a controlled time (10, 20, 30, 40, 50 or 60 min) in an oven. The water

was then transferred to a glass beaker. After cooling to room temperature, 10.0 mL of the

water sample was collected and subjected to the DSPE procedure, for quantification of

BPs. Migration kinetics of BPA and BPAF into the food simulants were evaluated by the

developed method after the different durations of storing.

Biphenol S (BPS) was a common coexisting substance in disposable plastic packaging.

Matrix effect of BPS on the determination of analytes migrated from disposable plastic

packaging (disposal plastic bag-Ⅱ) was studied using distilled water as food simulant at

100°C. Furthermore, in order to study matrix components for the analytes migration from

the disposable plastic packaging (disposal plastic bag-Ⅱ) according to Reg. (EU) No

10/2011 [19], four types of food simulants were chosen: distilled water (food simulant A),

3% acetic acid aqueous solution (food simulant B), 20% ethanol aqueous solution (V/V)

(food simulant C), and 50% ethanol aqueous solution (food simulant D1).

3. Result and discussion

3.1. Preparation and characterization of adsorbents

As shown in Fig. 1, the ZnO@CTAB composite was prepared by a typical

hydrothermal reaction using zinc acetate, urea, sodium citrate, and CTAB as raw

materials. The effect of CTAB quality on the adsorption performance of adsorbents was

investigated (Fig. S1). Notably, we observed that adding 0.6 g of CTAB could guarantee a

good adsorption performance of adsorbents. Further increase in the amount of CTAB led

9
to a slight decrease in BPs adsorption. It was also observed that the bare ZnO

microspheres showed no affinity for BPs, indicating that the effective adsorption of BPs

on ZnO@CTAB was due to the modification with CTAB. Thus, 0.6 g of CTAB was used

in the preparation of ZnO@CTAB and the received adsorbent was further used to extract

BPs from different matrixes.

The morphology of ZnO@CTAB was observed by TEM and SEM (Fig. S2). The

images showed that ZnO@CTAB formed aggregates of multiple particles with irregular

shapes, while the ZnO material was composed of 5–13 μm microspheres with a

hierarchical porous structure (Fig. S3). Compared with the ZnO microspheres,

ZnO@CTAB was smaller in size and more irregular in shape, which could be due to the

participation of CTAB in the formation of the ZnO crystal nuclei and the termination of

smaller particles to form larger microspheres.

The FT-IR spectra of ZnO and ZnO@CTAB are shown and compared in Fig. 2A. Both

showed the characteristic absorption peaks of Zn-O at 470 cm-1. Also, the characteristic

peak of the O-H bond was present at 3388 cm-1. For ZnO@CTAB, strong absorption

peaks appeared at 2921 and 2854 cm-1, which were attributed to the symmetric and

anti-symmetric stretching vibrations of C-H bonds. Additionally, the C-N stretching

vibration peak appeared at 1496 cm-1, confirming the existence of CTAB in the

ZnO@CTAB nonmaterial. The XRD pattern of ZnO@CTAB (Fig. 2B) shows that the

diffraction peaks matched well with the monoclinic phase of ZnO precursor (JCPDS No.

19-1458). All these results indicated that the adsorbent consisted of organic CTAB and

ZnO precursor.

In order to determine the thermal stability of adsorbents, the samples were analyzed by

10
thermogravimetric analysis. As shown in Fig. S4, the thermogravimetric curve of

ZnO@CTAB showed two-step degradation. The first stage of weight loss (0–200°C) was

similar to that for ZnO microspheres, which was mainly caused by the loss of surface

water molecules from ZnO. The second stage of weight loss (200–550°C) was attributed

to the decomposition of CTAB and the ZnO precursor to ZnO, CO 2 and water [20].

Fig. 2C presents the N2 adsorption/desorption isotherm and pore size distribution of

ZnO@CTAB. The adsorbent displayed a type-IV isotherm with an H3 hysteresis loop,

indicative of a mesoporous state. The Brunauer-Emmett-Teller (BET) surface area of

ZnO@CTAB was calculated to be 110.0 m2 g-1, which was much higher than that of the

individual ZnO microspheres reported in the literatures [21-23]. The

Barrett-Joyner-Halenda (BJH) curve showed that pore size centered at ~3.2 nm was due

to CTAB modification, and the pore size distributed from 5.4–35.0 nm was due to the

randomly dispersed pore sizes of ZnO@CTAB (Fig. 2C inset). Fig. 2D shows the

ζ-potentials of ZnO@CTAB and ZnO: after modification with CTAB, ZnO@CTAB was

positively charged in the pH range of 2.0–10.

3.2. Adsorption property of ZnO@CTAB

The adsorption of BPs at different concentrations by ZnO@CTAB is shown in Fig.

S5. As can be seen, the increase in concentration of BPs led to an obvious increase in

adsorption amount. The experimental data were then analyzed using the Langmuir and

Freundlich isotherm models. As shown in Table S1, the Langmuir model was more

suitable to describe BPs adsorption on ZnO@CTAB (with a higher R2 of 0.9599–0.9651),

indicating the monolayer adsorption on the homogeneous surface without interaction

11
among the adsorbate molecules. Table S2 shows the adsorption capacity (Qmax) of

different adsorbents for BPs. It was observed that the Qmax of ZnO@CTAB for BPA and

BPAF was 99.01 and 126.31 mg g-1, respectively, which was higher than that of

microplastics [24] , MgO-PAC crystals [25], modified carbon nanotubes [26] and

magnetic covalent organic framework nanocomposites [27]. Though the Qmax of

ZnO@CTAB was less than for magnetic iron oxide/graphene [28] and nanoporous carbon

derived from MOF materials [29], the preparation method was much simpler and more

environmentally friendly. Furthermore, worth noting is the fact that ZnO@CTAB showed

a very fast adsorption kinetics for BPA and BPAF with equilibrium time within 20 min

(Fig. S6); this indicated that the adsorbent was very favorable for use with the DSPE

method for the analysis of BPA and BPAF.

3.3. DSPE optimization

In order to obtain optimum DSPE conditions for quantitative analysis of BPs, the

experimental parameters of solution pH, ZnO@CTAB dosage, extraction time, type of

desorption solvent, and desorption time were investigated.

The extraction experiments were carried out over a pH range of 4.0–10.0 to ascertain

the role of initial pH on the recovery of BPs. As presented in Fig. 3A, better recovery was

observed from pH 7.0 to 9.0. These results could be ascribed to the surface charged

species of ZnO@CTAB and the nature of BPs at different pH values. BPA and BPAF are

very weak acids in water, with pKa values of 10.3 and 8.7, respectively [30]. BPs will be

present as the molecular form, which can be easily adsorbed, when the pH of the aqueous

phase is lower than the pKa values [31]. On the other hand, ZnO@CTAB is positively

12
charged over the pH range of 2.0–10.0 due to cationic CTAB on the surface. Therefore,

BPA and BPAF could easily be adsorbed on ZnO@CTAB via hydrophobic interaction

and cation-π electron interaction between BP benzene rings and the ammonium ion of

CTAB over the pH range of 4.0–10 [15, 32].. Given these results, and as the pH of pure

water is ~7.0, subsequent DSPE was performed without pH adjustment.

As shown in Fig. 3B, the recovery of BPs improved from 53.72–54.63% to 91.51–

94.11%, with increasing amount of adsorbent from 2 to 10 mg, above which the recovery

stabilized. Therefore, 10 mg of ZnO@CTAB was selected. Fig. 3C presents the recovery

of BPs for the developed ZnO@CTAB material at different adsorption times. The results

indicated that 10 min was long enough to obtain a satisfactory recovery.

Four desorption solvents, MeOH, MeCN, ethylacetate (EtOAc) and a mixed solution

of MeOH and MeCN (1:1; v/v), were tested for desorption of BPs from ZnO@CTAB.

From Fig. 3D, the best recovery was obtained with MeOH/MeCN. Furthermore, as the

desorption time exceeded above 10 min, the extraction efficiency was relative constant

(Fig. S7). Therefore, 10 min was selected for desorption of BPs on ZnO@CTAB using

the mix solution of MeOH and MeCN as desorption solvent.

3.4. Method validation

The proposed method for BP analysis based on ZnO@CTAB coupled with

UPLC/Q-TOF-MS was assessed under the optimized conditions, as determined above. As

shown in Table 1, the two BPs showed good linearity (r = 0.9993 and 0.9989) in the

concentration range of 0.1–100 μg L-1. The expanded uncertainty of linear equations for

BPA and BPAF were 8.13% and 6.41%, respectively, with a coverage factor of 2 at 95%

13
confidence level (n=15). The limit of detections (LODs, S/N = 3) of BPA and BPAF

were 0.027 and 0.030 μg L-1, respectively, and the limit of quantifications (LOQs,

S/N=10) were 0.08 and 0.10 μg L-1, respectively. As shown in Tables S3 and S4, the

relative extraction recovery of the analytes was measured to be between 97.63–109.33%.

Furthermore, intra-day precisions were calculated at three concentration levels of 0.5, 1

and 10 μg L-1 in a single day, with the RSDs ranging from 0.71–3.54%. The inter-day

precisions were calculated at 0.5, 1 and 10 μg L-1 on three consecutive days with the

RSDs ranging from 1.67–6.23%.

Table 2 shows a comparison of the proposed method with recent adsorbent-based SPE

methodologies for analysis of trace BPs [33-39]. As can be seen, the adsorbent dosage

used in the developed method was 10 mg which was lower than reported approaches [33,

34, 36, 38, 39]. The DSPE time (including adsorption and desorption time) spent in the

developed method was acceptable compared with that of reported in other references.

Furthermore, the method exhibited comparable or even better sensitivity compared with

the previously reports except for SPE/HPLC-MS/MS which used a much sensitive

detector [39]. At the same time, the extraction efficiency in this study were higher than

that of reported in the references. What’s more, the preparation method of ZnO@CTAB

was much easier and more environmentally friendly, which is crucial for the DSPE

adsorbents real application.

3.5. Migration tests and matrix effect

Due to the high efficiency of food delivery, most food is delivered within one hour.

Therefore, the release of BPs from disposable plastic bags and disposable plastic

14
container were evaluated from 10 to 60 min at 100°C. As shown in Fig. 4, BPAF

migrated from both disposable plastic bags and the levels of BPAF increased

dramatically when the storage time increased from 0 to 40 min. The migrated

concentration of BPAF was found to be 0.42 μg L-1 and 0.86 μg L-1 for disposable plastic

bag-Ⅰ and disposable plastic bag-Ⅱ, respectively. For disposable plastic container, 0.12

μg L-1 of BPA was observed within 10 min, which increased to 0.49 μg L-1 after 40 min.

Fig. 5 shows the typical MS spectra and chromatograms of the migrated BPA and BPAF

from these disposable plastic packing at 40 min. The significant migration of BPA and

BPAF into the hot liquid food simulant revealed that the use of plastic materials in

takeaway food poses in a high risk of human exposure.

The influence of spiked BPS (1 and 5 μg L-1) on the determination of analytes migrated

from disposable plastic bag-Ⅱ in distilled water was examined. The results showed that

the migration levels of BPAF were 0.859 (spiked with BPS at 1 μg L-1), and 0.863 μg L-1

(spiked with BPS at 5 μg L-1), respectively, which were consistent with the results

without BPS spiked. These results indicated that the coexisting BPS could not affect

DSPE of the target analytes on ZnO@CTAB in real matrix.

BP migration from the disposable plastic materials was not only dependent on the

storage conditions (time and temperature), but also on the food components. Fig. 6 shows

the migration levels of BPAF to different liquid food simulants during the storage of

disposable plastic bag-Ⅱ at 100 °C. The results showed that the migration levels of BPAF

after 40 min in food simulants A, B, C, and D1 were 0.86, 1.07, 1.46 and 1.57 μg L-1,

respectively. BPAF levels in distilled water were lower than those measured in acidic

food, ethanol solution and fatty food simulants. The quantity of BPAF that migrated into

15
simulant D1 was about two times that found into simulant A. The data confirmed a

significant release of BPAF levels depending on the polarity of the liquid food: the lower

the polarity, the greater the quantity of BPAF that migrated into the food. To test the

precision of the developed DSPE method, relative recovery values were obtained in

different food simulants spiked with BPAF (1 and 5 μg L-1). The acceptable RSDs (1.37–

5.64) and relative recovery values (95.12–109.40%) for different food stimulants

containing released BPAF supported the practicability of the newly developed method for

real sample analysis.

4. Conclusion

A CTAB-intercalated mesoporous ZnO composite was prepared by a facile synthesis

method and a new analytical method based on the as-received adsorbent was developed

for the quantification of BPs in water samples. The applicability of the DSPE method for

the migration of BPs from plastics materials to liquid food simulants was also examined.

BPAF was found to migrate from both disposable plastic bags, yielding BP

concentrations of 0.42 μg L-1 and 0.86 μg L-1, respectively. BPA was found to migrate

from disposal plastic container to food simulates to give a concentration of 0.49 μg L-1

when the food simulates were kept at 100°C for 40 min. In addition, food components

played a major role in favoring migration. Surprisingly, the quantity of BPAF that

migrated into fatty food simulants was about two times that found into distilled water.

These results indicated that exposure to BPA and BPAF from plastic packaging materials

should not be ignored.

16
Acknowledgment: The present work is supported by the projects of the National Natural

Science Foundation of China (21605056), Natural Science Foundation of Jilin Province

(20180101292JC), and the Science and Technology Development of Jilin province

(20190303116S). Financial support from the Key Laboratory of Fine Chemicals of Jilin

Province is also acknowledged.

Declaration of interests

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

Appendix A. Supplementary data

References

[1] A. Gallego-Schmid, J.M.F. Mendoza, A. Azapagic, Environmental impacts of takeaway food containers,
J. Clean.Prod. 211 (2019) 417-427.
[2] D. Wu, C. Huang, X. Jiao, Z. Ding, S. Zhang, Y. Miao, L. Huo, Bisphenol AF compromises blood-testis
barrier integrity and sperm quality in mice, Chemosphere 237 (2019) 124410.
[3] B. Jia, T. Shi, Z. Li, S. Shan, P. Ji, Z. Li, Toxicological effects of bisphenol A exposure-induced cancer
cells migration via activating directly integrin β1, Chemosphere 220 (2019) 783-792.
[4] Y. Liu, Z. Yan, Q. Zhang, N. Song, J. Cheng, O.L. Torres, J. Chen, S. Zhang, R. Guo, Urinary levels,
composition profile and cumulative risk of bisphenols in preschool-aged children from Nanjing suburb,
China, Ecotox.Environ. Safe. 172 (2019) 444-450.
[5] G. Russo, F. Barbato, D.G. Mita, L. Grumetto, Occurrence of Bisphenol A and its analogues in some
foodstuff marketed in Europe, Food Chem. Toxicol. 131 (2019) 110575.
[6] A. Ostovan, M. Ghaedi, M. Arabi, Q. Yang, J. Li, L. Chen, Hydrophilic Multitemplate Molecularly
Imprinted Biopolymers Based on a Green Synthesis Strategy for Determination of B-Family Vitamins,
ACS Appl. Mater. Interfaces 10 (2018) 4140-4150.
[7] N. Li, J. Du, D. Wu, J. Liu, N. Li, Z. Sun, G. Li, Y. Wu, Recent advances in facile synthesis and
applications of covalent organic framework materials as superior adsorbents in sample pretreatment, TrAC
Trends Anal. Chem. 108 (2018) 154-166.
[8] A.R. Bagheri, M. Arabi, M. Ghaedi, A. Ostovan, X. Wang, J. Li, L. Chen, Dummy molecularly
imprinted polymers based on a green synthesis strategy for magnetic solid-phase extraction of acrylamide
in food samples, Talanta 195 (2019) 390-400.
[9] Y. Wen, L. Chen, J. Li, D. Liu, L. Chen, Recent advances in solid-phase sorbents for sample
preparation prior to chromatographic analysis, TrAC Trends Anal. Chem. 59 (2014) 26-41.
[10] J. Li, J. Liu, W. Lu, F. Gao, L. Wang, J. Ma, H. Liu, C. Liao, L. Chen, Speciation analysis of mercury
by dispersive solid-phase extraction coupled with capillary electrophoresis, Electrophoresis. 39 (2018)
1763-1770.
[11] Y. Wen, Z. Niu, Y. Ma, J. Ma, L. Chen, Graphene oxide-based microspheres for the dispersive
solid-phase extraction of non-steroidal estrogens from water samples, J. Chromatogr. A 1368 (2014) 18-25.
[12] M.N. Zafar, Q. Dar, F. Nawaz, M.N. Zafar, M. Iqbal, M.F. Nazar, Effective adsorptive removal of azo

17
dyes over spherical ZnO nanoparticles, J. Mater. Res. Technol. 8 (2019) 713-725.
[13] N. Dhiman, N. Sharma, Removal of pharmaceutical drugs from binary mixtures by use of ZnO
nanoparticles: (Competitive adsorption of drugs), Environ. Technol. Inno. 15 (2019) 100392.
[14] K. Chaiyasing, B. Liawruangrath, S. Natakankitkul, S. Satienperakul, N. Rannurags, P. Norfun, S.
Liawruangrath, Sequential injection analysis for the determination of fluoroquinolone antibacterial drug
residues by using eosin Y as complexing agent, Spectrochim. Acta. Part A (2018) 107-114.
[15] L. Wang, Y. Cao, Adsorption behavior of phenanthrene on CTAB-modified polystyrene microspheres,
Colloid. Surface. A 553 (2018) 689-694.
[16] X. Zhang, Y. Yang, L. Song, J. Chen, Y. Yang, Y. Wang, Enhanced adsorption performance of
gaseous toluene on defective UiO-66 metal organic framework: Equilibrium and kinetic studies, J. Hazard.
Mater. 365 (2019) 597-605.
[17] J. Guo, S. Chen, L. Liu, B. Li, P. Yang, L. Zhang, Y. Feng, Adsorption of dye from wastewater using
chitosan–CTAB modified bentonites, J. Colloid Interf. Sci 382 (2012) 61-66.
[18] L. Wang, Y. Cao, Adsorption behavior of phenanthrene on CTAB-modified polystyrene microspheres,
Colloid Surface A 553 (2018) 689-694.
[19] E. Commission, Commission Regulation (EU) No 10/2011 of 14 January 2011 on plastic materials and
articles intended to come into contact with food, Off J. Eur. Union 12 (2011) 1-89.
[20] M. Khoeini, A. Najafi, H. Rastegar, M. Amani, Improvement of hollow mesoporous silica
nanoparticles synthesis by hard-templating method via CTAB surfactant, Ceram. Int. 45 (2019)
12700-12707.
[21] S. Bera, H. Khan, I. Biswas, S. Jana, Polyaniline hybridized surface defective ZnO nanorods with
long-term stable photoelectrochemical activity, Appl. Surf. Sci. 383 (2016) 165-176.
[22] F.A. Lucca Sánchez, A.S. Takimi, F. Severo Rodembusch, C. Pérez Bergmann, Photocatalytic activity
of nanoneedles, nanospheres, and polyhedral shaped ZnO powders in organic dye degradation processes, J.
Alloy. Compd. 572 (2013) 68-73.
[23] A. Lei, B. Qu, W. Zhou, Y. Wang, Q. Zhang, B. Zou, Facile synthesis and enhanced photocatalytic
activity of hierarchical porous ZnO microspheres, Mater. Lett. 66 (2012) 72-75.
[24] P. Wu, Z. Cai, H. Jin, Y. Tang, Adsorption mechanisms of five bisphenol analogues on PVC
microplastics, Sci. Total. Environ. 650 (2019) 671-678.
[25] B. Kamarehie, S.M.S. Tizabi, R. Heydari, A. Jafari, M. Ghaderpoori, M.A. Karami, A. Ghaderpoury,
Data on the bisphenol A adsorption from aqueous solutions on PAC and MgO~PAC crystals, Data in Brief
21 (2018) 746-752.
[26] C. Kuo, Comparison with as-grown and microwave modified carbon nanotubes to removal aqueous
bisphenol A, Desalination. 249 (2009) 976-982.
[27] L. Chen, Y. He, Z. Lei, C. Gao, Q. Xie, P. Tong, Z. Lin, Preparation of core-shell structured magnetic
covalent organic framework nanocomposites for magnetic solid-phase extraction of bisphenols from human
serum sample, Talanta 181 (2018) 296-304.
[28] L.D. Quan, N.H. Dang, T.H. Tu, V.N. Phuong Linh, L.T. Mong Thy, H.M. Nam, M.T. Phong, N.H.
Hieu, Preparation of magnetic iron oxide/graphene aerogel nanocomposites for removal of bisphenol A
from water, Synthetic. Met. 255 (2019) 116106.
[29] S. Ma, X. Wang, H. Duan, J. Wang, H. Zhan, Z. Zhang, A nanoporous carbon derived from bimetallic
organic-framework for magnetic solid-phase extraction of bisphenol analogs, Talanta 202 (2019) 479-485.
[30] X. Sun, J. Wang, Y. Li, J. Jin, B. Zhang, S.M. Shah, X. Wang, J. Chen, Highly selective dummy
molecularly imprinted polymer as a solid-phase extraction sorbent for five bisphenols in tap and river water,
J. Chromatogr. A 1343 (2014) 33-41.
[31] Q. Zhou, M. Lei, Y. Wu, X. Zhou, H. Wang, Y. Sun, X. Sheng, Y. Tong, Magnetic solid phase
extraction of bisphenol A, phenol and hydroquinone from water samples by magnetic and thermo
dual-responsive core-shell nanomaterial, Chemosphere 238 (2020) 124621.
[32] H. Jiang, Y. Lin, N. Li, Z. Wang, M. Liu, R. Zhao, J. Lin, Application of magnetic N-doped carbon
nanotubes in solid-phase extraction of trace bisphenols from fruit juices, Food Chem. 269 (2018) 413-418.
[33] X. Hu, X. Wu, F. Yang, Q. Wang, C. He, S. Liu, Novel surface dummy molecularly imprinted silica as
sorbent for solid-phase extraction of bisphenol A from water samples, Talanta 148 (2016) 29-36.
[34] X. Wang, C. Deng, Preparation of magnetic graphene @polydopamine @Zr-MOF material for the
extraction and analysis of bisphenols in water samples, Talanta 144 (2015) 1329-1335.
[35] N. Li, J. Chen, Y. Shi, Magnetic nitrogen-doped reduced graphene oxide as a novel magnetic
solid-phase extraction adsorbent for the separation of bisphenol endocrine disruptors in carbonated

18
beverages, Talanta 201 (2019) 194-203.
[36] M. Gao, W. Liu, X. Wang, Y. Li, P. Zhou, L. Shi, B. Ye, R.A. Dahlgren, X. Wang,
Hydrogen-bonding-induced efficient dispersive solid phase extraction of bisphenols and their derivatives in
environmental waters using surface amino-functionalized MIL-101(Fe), Microchem. J. 145 (2019)
1151-1161.
[37] Q. Yao, Y. Feng, C. Tan, S. Xia, L. Zhao, S. Wang, Y. Wang, X. Chen, An on-line solid-phase
extraction disc packed with a phytic acid induced 3D graphene-based foam for the sensitive HPLC-PDA
determination of bisphenol A migration in disposable syringes, Talanta 179 (2018) 153-158.
[38] J. Zhang, Y. Chen, W. Wu, Z. Wang, Y. Chu, X. Chen, Hollow porous dummy molecularly imprinted
polymer as a sorbent of solid-phase extraction combined with accelerated solvent extraction for
determination of eight bisphenols in plastic products, Microchem. J. 145 (2019) 1176-1184.
[39] X. Sun, J. Peng, M. Wang, J. Wang, C. Tang, L. Yang, H. Lei, F. Li, X. Wang, J. Chen, Determination
of nine bisphenols in sewage and sludge using dummy molecularly imprinted solid-phase extraction
coupled with liquid chromatography tandem mass spectrometry, J. Chromatogr. A 1552 (2018) 10-16.

Captions:

Fig. 1. Schematic of preparation of ZnO@CTAB and its application in the extraction of

BP migration from disposable plastic packaging materials.

Fig. 2. The FTIR spectra of ZnO@CTAB and ZnO (A); XRD partten of ZnO@CTAB

(B); Nitrogen adsorption-desorption isotherm and pore size distribution of ZnO@CTAB

(C); ζ-potentials of ZnO@CTAB and ZnO (D).

Fig. 3. Effect of the pH (A), adsorbent amount (B), extraction time (C), and desorption

solvents (D) on the extraction of BPA and BPAF.

Fig. 4. Effect of storage time on the migration of BPA and BPAF from disposable plastic

19
bag-Ⅰ, disposable plastic bag-Ⅱ and disposable plastic container.

Fig. 5. Q-TOF mass spectra and chromatograms obtained for disposable plastic bag-Ⅰ

(A), disposable plastic bag-Ⅱ (B) and disposal plastic container (C). (On the left are the

Q-TOF mass spectra and on the right are UPLC chromatograms).

Fig. 6. Migration of BPAF from disposable plastic bag-Ⅱ into different matrixes.

Fig. 1

20
Fig. 2

Fig. 3

21
Fig. 4

22
Fig. 5

23
Fig. 6

24
Table 1

Calibration curves, LODs, LOQs and r of the developed method (n=3).

Analytes Quantification equations Linearity range r LOD LOQ

(μg L-1) (μg L-1) (μg L-1)

BPA y=1183x-55.542 0.1–100 0.9993 0.027 0.08

BPAF y= 1221.9x+22.927 0.1–100 0.9989 0.030 0.10

Table 2

Comparison of the proposed method with recent adsorbent-based SPE methodologies for

25
analysis of trace BPs

Adsorbent Quality Instrumental DSPE Recovery LODs Sample Ref.

method time (%) (μg L-1)

SiO2@MIP 40 mg HPLC-UV – 97.3–106.0 0.3 Water [33]

Magnetic MOF 30 mg HPLC-UV 25 min 64.8–92.8 0.1-1 Water [34]

Fe3O4@N-RGO 3 mg HPLC-DAD 8 min 86.5–101.5 0.1–0.2 Beverages [35]

NH2-MIL-101(Fe) 30 mg HPLC 3 min 93.1–105.8 0.1 Water [36]

PAGF 2 mg HPLC-PDA – 73–117 0.03 Water [37]

HPDMIP 100 mg UPLC-MS/MS – 50.8–138.3 0.02–0.6 Plastic [38]

sample

MIPs 200 mg UPLC-MS/MS 105 min 43.6–101 0.0007–16.3 ng L-1 Sewage [39]

ZnO@CTAB 10 mg UPLC-TOF-MS 20 min 97.6–109.3 0.027–0.030 Water This work

26

You might also like