Sei sulla pagina 1di 29

Authors Accepted Manuscript

Controlled synthesis of hollow porous carbon


spheres for enrichment and simultaneous
determination of nine bisphenols from real samples

Zhenzhen Zhang, Jing Zhang, Yang Wang, Yao


Tong, Lei Zhang
www.elsevier.com/locate/talanta

PII: S0039-9140(17)30251-5
DOI: http://dx.doi.org/10.1016/j.talanta.2017.02.037
Reference: TAL17313
To appear in: Talanta
Received date: 11 November 2016
Revised date: 10 February 2017
Accepted date: 17 February 2017
Cite this article as: Zhenzhen Zhang, Jing Zhang, Yang Wang, Yao Tong and Lei
Zhang, Controlled synthesis of hollow porous carbon spheres for enrichment and
simultaneous determination of nine bisphenols from real samples, Talanta,
http://dx.doi.org/10.1016/j.talanta.2017.02.037
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
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.
Controlled synthesis of hollow porous carbon spheres for enrichment

and simultaneous determination of nine bisphenols from real samples

Zhenzhen Zhang, Jing Zhang, Yang Wang, Yao Tong, Lei Zhang

College of Chemistry, Liaoning University, 66 Chongshan Middle Road, Shenyang,

Liaoning, 110036, Peoples Republic of China

ABSTRACT

An extended one-step Stber method was utilized for the preparation of

core@shell spheres, which was made up of a thin layer of resorcinol-formaldehyde

(RF) and a silica core. After the carbonization and template-removal process, hollow

porous carbon spheres (HPCSs) were synthesized. The structure of HPCSs was

characterized by scanning/transmission electron microscopy and N2

adsorption/desorption isotherms, Fourier transform infrared spectroscopy, energy

dispersive X-ray spectrometer and Raman spectroscopy. The results showed that the

silica cores were removed successfully and the HPCSs were in good sphere shape

with uniform size, high surface areas as well as pores hollow framework structure.

Compared with MWCNTs and 3D-graphene, HPCSs exhibited superior extraction

ability for nine bisphenols (BPs). HPCSs showed extremely outstanding extraction

efficiency for BPs as well as high adsorption capacity due to its hollow porous

Corresponding author.
Lei Zhang, Ph.D.,
College of Chemistry, Liaoning University,
110036, Peoples Republic of China
Tel.: +86 24 62207809; Fax: +86 24 62202380.

E-mail address:
zhanglei63@126.com (L. Zhang).
1
structure. It was applied as sorbent for the enrichment of nine BPs from

environmental water samples and soft drinks prior to high performance liquid

chromatographic (HPLC) analysis, obtaining recoveries ranged from 89.6% to

111.5% and the relative standard deviations (RSD) from 0.8% to 6.3%, and the limits

of quantification (LOQs) were in the range of 0.183-1.763 g L-1.

Keywords: Hollow porous carbon spheres; monodisperse; Bisphenols; solid-phase

extraction

1. INTRODUCTION

Sample preparation has been regarded as a critical step in analytical laboratory

when performing multiple analyses, and it is also the key to accurate analysis [1]. To

date, a variety of sample preparation methods have been developed for the isolation

and preconcentration of target analytes from different samples. Solid phase extraction

(SPE) is one of the most popular techniques due to its high enrichment performance,

low solvent cost and simple operation process [2]. In general, the sorbent plays a vital

role for the efficient extraction of the analytes in SPE [3]. Therefore, the exploration

of new sorbents has been the focus of attention in this research area. The application

of carbon-based materials including fullerene [4], carbon nanotubes [5], carbon

nanofibers [6], carbon nanocones/disks [7], graphene [8] and 3D-graphene [9], has

attracted significant attention in SPE for separation and extraction of various

compounds due to their high adsorption capacity and large surface area [10]. Among

these materials, MWCNTs and graphene are particularly suited for sorption of

aromatic compounds because of interaction and Van der Waals force [11].

2
However, the use of MWCNTs as sorbent alone can be problematic due to their

extreme hydrophobicity and high Van der Waals interaction forces along the length

axis, resulting in aggregation and deposition in water [12], and disadvantages related

to tedious synthetic procedures, high manufacturing cost and low yields have impeded

large scale application [13]. Besides, recent studies have shown that the restacking or

aggregation between individual graphene sheets is usually unavoidable due to the

strong stacking, hydrophobic interactions, and van der Waals forces, resulting in

decrease of surface area, which is undesirable for its applications [14].

Nowadays, functional porous carbon spheres have become an appealing topic in

the carbon community [15] because their unique structure and functional behavior

(such as greater pore accessibility, higher surface area, lower density, faster molecular

diffusion/transfer and more controllable inner pore volume [16]) make them quite

useful in adsorption [17], drug delivery [18], energy storage [19] and catalysis [20].

The aim of this work is to design and develop a new kind of adsorption material,

which can gather all the advantages of graphene, MWCNTs and carbon nanofibers,

and lead to the excellent performance in practical applications. For example, the

integrated material will be able to serve as powerful solid-phase extraction sorbents

for sample pretreatment technique in analytical chemistry.

In the current work, an extended one-step Stber method was utilized for the

preparation of a novel SiO2@resorcinol-formaldehyde (SiO2@RF) spheres. After the

carbonization and template-removal process, monodisperse hollow porous carbon

spheres (HPCSs) with high dispersity were synthesized. The prepared HPCSs possess

3
three-dimensional nanostructure, high specific surface area and abundant mesopores.

Besides, the carbon spheres can form a strong - interaction with the benzene

ring-based compounds due to the presence of graphitic carbons [21]. Therefore they

could be good sorbents for the extraction of aromatic compounds. To investigate the

extraction ability of HPCSs, nine bisphenols (BPs) including bisphenol S (BPS),

bisphenol F (BPF), 4,4'-Thiodiphenol (TDP), bisphenol A (BPA), bisphenol AF

(BPAF), bisphenol AP (BPAP), bisphenol C (BPC), tetrachlorobisphenol A (TCBPA)

and tetrabromobisphenol A (TBBPA) were selected as the model compounds (Fig.

S-1, supplementary material). BPs are a group of chemicals with two hydroxyphenyl

functionalities and include several structural analogs [22], which are widely used as

additives to improve the properties of plastics [23]. It was reported that BPs had

estrogenic activity, and were potentially carcinogenic and mutagenic to animals and

humans [24, 25]. In recent years, there has been a growing concern worldwide about

the appearance of BPs in human fluid, drinking water and environmental water

samples [26], so reliable analytical methods for simultaneous determination of BPs

were required. Herein, HPCSs were controlled synthesized and characterized as

solid-phase extraction sorbents, which, to the best of our knowledge, has not been

used in this field. A HPCSs-based SPE method was established and applied for the

enrichment and determination of trace levels of nine BPs in large-volume

environment water and juice samples prior to high-performance liquid

chromatography-diode array detection (HPLC-DAD) analysis.

2. EXPERIMENTAL

4
2.1 Reagents and materials

All chemicals were analytical grade, unless otherwise noted. Graphite powder

(natural flake graphite, 325 mesh) was purchased from Alfa Aesar. MWCNTs were

purchased from Chengdu Organic Chemicals Co., Ltd, China. BPS was purchased

from TCI (Shanghai) Development Co., Ltd.; BPF, BPA, BPAF, BPAP, BPC and

TCBPA were purchased from HEOWNS Biochemical technology Co., Ltd. (Tianjin,

China); TDP and TBBPA were purchased from Shanghai Aladdin industrial Co., Ltd..

Deionized water purified using a Sartorius Arium 611 system (Sartorius, Gttingen,

Germany) was used throughout the experiment. HPLC grade methanol was purchased

from Fisher Corporation (Pittsburgh, PA, USA). HPLC grade ultrapure water (18.2

M cm-1 resistivity) was obtained from a MilliQ water purification system (Millipore,

Milford, MA, USA)

The stock solutions (1.0 g L-1) containing the analytes was prepared by

dissolving appropriate amount of them individually in methanol and stored at 4 C

under dark conditions. The working solutions were obtained daily by appropriately

diluting the stock solutions with deionized water.

2.2 Preparation of HPCSs and 3D-G

Based on the method previously reported [27], only the amounts of raw materials

were optimized in terms of the production of HPCSs. Briefly, in a typical synthesis

process, 1.6 mL of ammonia aqueous solution (25%~28%, w/w) was added to a

mixture of 5.5 mL deionized water and 37.8 mL absolute ethanol and stirred for 30

minutes at room temperature. Subsequently, tetrathoxysilane (TEOS) (3.0 mL),

5
resorcinol (0.42 g) and formaldehyde solution (0.62 mL, 37%, w/w) were added to the

above solution at intervals of 10 minutes. Then the mixture was vigorously stirred at

30 C for 24 h and transferred to a 100 mL Teflon-lined autoclave maintained under at

100 C for 24 h. Solid SiO2@RF products were collected by centrifugation and

washed several times with absolute ethanol and deionized water, and then dried at

60 C overnight. After oven-drying, the solid samples were heated at 750 C for 2 h

under nitrogen gas flow in a tube furnace. Next, the SiO2 template was dissolved in

10% HF (w/w) solution overnight and washed several times with deionized water to

be neutral. Finally, the HPCSs were obtained after dried at 60 C for 12 h. The

schematic synthesis process of HPCSs was presented in Fig. 1. To investigate the

batch-to-batch reproducibility, the extraction efficiency of three independently

prepared batches of HPCSs was evaluated and the data were all between 97% and

99%, showing good reproducibility. So in the following experiments, all batches of

HPCSs were mixed together for use.

Graphene oxide (GO) was synthesized from graphite powder using a modified

Hummers method [28]. The 3D-G was fabricated by a facile one-step hydrothermal

method. Firstly, 120.0 mg of GO was added into 60.0 mL of ionized water and the

suspension was ultrasonic dispersed for 1 h. Next, the GO dispersion was then poured

into a 100 mL Teflon-lined autoclave and heated at 180 oC for 12 h. Finally, the 3D-G

was obtained by freezing dry.

2.3 Characterization of the HPCSs

The BrunauerEmmettTeller (BET) specic surface area and pore size

6
distribution of HPCSs were measured using N2 adsorption/desorption isotherm on an

ASIQ-C automated gas sorption analyzer (Quantachrome, USA). TEM (Transmission

Electron Microscopy) images were obtained using a JEM-2100HR microscope (JEOL,

Japan) operated at 200 kV. SEM (Scanning Electron Microscopy) images were

obtained using HITACHI SU8000 scanning electron microcopy at 10.0 kV. Fourier

transformed infrared (FTIR) spectra (4000-400 cm-1) of SiO2@C and HPCSs in KBr

were recorded using the Nicolet FT-IR 5700. The composition of the samples was

analyzed by using an energy dispersive X-ray spectrometer (EDX) attached to SEM.

Raman spectrum was measured using Laser Confocal Micro-Raman Spectroscope

(LabRAM XploRA, 140 HORIBA JOBIN YVON S.A.S, France).

2.4 Apparatus and operating conditions for HPLC

HPLC analysis was performed with an Agilent 1100 HPLC system (Palo Alto,

CA, USA) equipped with an automatic sampler and diode array detector (DAD).

Chromatographic separation of target analytes was performed on a ZOABAX Eclipse

SB-C18 (150 mm4.6 mm, 5 m) column (Agilent, Palo Alto, CA, USA) and the

injection volume was 20.0 L. A gradient elution was adopted by combining solvent

A (methanol) and solvent B (water) as follows: 40% - 70% A (10 min), 70% - 100% A

(10 min), 100% - 40% A (5 min). The column temperature was kept at 30 C, the flow

rate was 1.0 mL min-1, and the detection wavelength was set at 226 nm.

2.5 Adsorption capacity experiments

All of the adsorption experiments were performed as follows: 5.0 mg of the

as-prepared HPCSs was added into 200.0 mL solutions of each BPs at the

7
concentration of 50.0 mg L-1 and the suspensions were stirred vigorously with a speed

of 1000 rpm throughout the experiments. After the adsorption reached equilibrium,

the sorbent was separated by centrifugation for 10.0 min, and the supernatant was

analyzed by HPLC.

The adsorption capacity of HPCSs for BPs was calculated using the following

equation:
C Ce V
Q 0
m
where Q (mg g-1) is the adsorption capacity, C0 (mg L-1) is the initial concentration of

the solution, Ce (mg L-1) is the equilibrium concentration of the solution, V (mL) is the

volume of the solution, and m (mg) is the mass of HPCSs.

2.6 Solid phase extraction procedure

Briefly, an amount of 5.0 mg of HPCSs was added to 10.0 mL of the working

standard solution (or 200.0 mL of spiked sample). After performing the extraction by

ultrasonic for 30 s at room temperature, the entire volume of the solution was passed

through a 0.45m nylon filter (14 mm i.d.) using a syringe. Then, the analytes were

eluted by passing 4.0 mL of methanol through the filter, and the eluate was evaporated

to dryness under a gentle N2 stream. The residue was redissolved in 0.5 mL of

methanol. The resulting solution was referred as analytical solution and stored at 4 C

for HPLC analysis.

2.7 Sample treatment

Sea water, river water and grape juice were collected as the real samples. Sea

water samples were collected from Bohai Sea. River water was collected from Xinkai

8
River (Shenyang, China). Grape juice was purchased from a local supermarket

(Shenyang, China). The samples were first centrifuged at 10000 rpm for 10 min and

the supernatant was collected. After filtering through 0.22 m filter membrane, sea

water and river water (for grape juice, an extra diluting process with equal volume of

deionized water was added) were stored in brown glass bottles at 4 C prior to

analysis. The pH of treated sea water, river water and grape juice were approximately

7.7, 6.7 and 3.2, respectively.

3. RESULTS AND DISCUSSION

3.1 Characterization of MWCNTs, 3D-G and HPCSs

The morphology of HPCSs, 3D-G and MWCNTs was investigated by SEM and

TEM. From Fig. 2A-C, it was observed that the synthesized HPCSs were relatively

uniform and monodisperse spheres with average diameter of about 400 nm. In general,

the monodisperse property could help sorbents disperse more easily in sample

solution and would further improve the extraction efficiency. The photographs of the

dispersion status of HPCSs (sonication time: 30 s) with different time from 1day to 4

days were shown in the inset of Fig. 2A. As seen from the photographs, HPCSs

displayed very good dispersity in solution, and the dispersion could be stable for

almost three days. In Fig. 2C, the hollow structure was confirmed by TEM, and the

wall thickness was about 20 nm. From Fig. S-2A (supplementary material), it could be

seen that 3D-G was of three dimensional architecture with a plenty of pores. As seen

from Fig. S-2B (supplementary material), MWCNTs were highly entangled with the

diameter of several tens of nanometers.

9
To compare the specific areas of home-made HPCSs with MWCNTs and 3D-G,

the BET specific surface areas and pore size distribution of the MWCNTs, 3D-G and

HPCSs were characterized using N2 adsorption/desorption isotherms. As seen in

Fig.3A, the BET specific surface areas of MWCNTs, 3D-G and HPCSs were 343.603,

312.889 and 571.278 m2 g-1, respectively. The higher surface area of HPCSs could be

attributed to its hollow and porous structure. The pore volume for MWCNTs, 3D-G

and HPCSs were 0.706, 3.300 and 1.710 cm3 g-1, respectively. The pore diameters

(maximum of the pore size distribution curve obtained from the isotherms adsorption

branch) for MWCNTs, 3D-G and HPCSs were 30.202 nm, 3.307nm and 3.494 nm,

respectively, which could all be attributed to mesopores (between 2 to 50 nm). The

high specific surface areas and high pore volumes would help improve the extraction

efficiency of the adsorption materials.

FT-IR spectra and EDS spectra were used to verify that the templates were

removed by HF. From Fig. 3B, it was observed the characteristic absorptions of

Si-O-Si at around 1095 cm-1 disappeared after the treating process by HF. The EDS

results were in accordance with that of FT-IR spectra. From Fig. 3C, it could be seen

that the three constituent elements C, O, and Si could all be detectable in SiO2@C,

while after the etching process, the peak of Si element disappeared. All the results

proved the successful removal of SiO2.

Fig. 3D displayed the Raman spectrum of HPCSs. As could be seen, the

spectrum showed the existence of the D and G bands. The G band corresponds to the

first-order scattering of the E2g mode observed for sp2 carbon domains, while the

10
pronounced D band is caused by structural defects or edges that break the symmetry

and selection rule [29]. The existence of sp2 carbon domains made it possible to form

interactions between HPCSs and benzene rings in BPs.

3.2 Comparison of extraction ability of HPCSs with MWCNTs and 3D-G

To demonstrate the superior extraction ability of HPCSs, the extraction ability of

HPCSs was compared with traditional CMs including MWCNTs and 3D-G from two

aspects: extraction efficiency and adsorption capability.

The extraction efficiency of different CMs was evaluated by applying 5.0 mg of

MWCNTs, 3D-G and HPCSs to extract 10.0 mL 2.0 mg L-1 solutions of nine BPs.

From Fig. 4A, it was observed that HPCSs displayed the highest extraction efficiency

for BPs.

To further testify the superior extraction ability of HPCSs, the adsorption

capacity experiment was investigated. The adsorption capabitilies of MWCNTs, 3D-G

and HPCSs for 50.0 mg L-1 BPs (volume: 200.0 mL) were shown in Fig. 4B. It could

be seen that HPCSs showed the highest adsorption capabilities for BPs, indicating the

outstanding extraction ability of HPCSs for BPs. The stacking interactions

between the bulk system on the HPCSs surface and BPs may be the main adsorption

mechanism. Besides, the hydrogen bond interaction may be formed between the OH

groups of BPs and the O-containing groups of HPCSs, and the benzene ring on

HPCSs surface may also act as a hydrogen bond donor and form hydrogen bonds with

OH functional groups on BPs [30].

3.3 Optimization of the extraction conditions

11
To achieve the best extraction efficiency of the targets with HPCSs based SPE

process, extraction conditions were optimized by analyzing spiked samples

(concentration: 0.5 mg L-1; volume: 10.0 mL). Several experimental parameters, such

as dosage of HPCSs, sonication time, pH of the solution, ionic strength and elution

conditions were investigated.

3.3.1 Effect of HPCSs amount dispersed in sample solution

Different dosages of the HPCSs in a range of 1.0-7.0 mg were dispersed in

sample solutions to extract BPs. The results (Fig. S-3A, supplementary material)

indicated that 5.0 mg HPCSs was enough for the quantitative the extraction of nine

BPs, and the extraction efficiency was over 96.7%. Therefore, 5.0 mg HPCSs was

selected to be applied to SPE process for the following experiments. High extraction

efficiency of BPs from aqueous phase may be attributed to the large surface area and

high dispersion of HPCSs.

3.3.2 Effect of extraction time

As an equilibrium-based procedure, the extraction time in SPE is one of the key

factors influencing the extraction efficiency. This parameter was investigated by

varying sonication time from 5 to 40 s in the experiment. The plot of the extraction

time versus extraction efficiency of BPs was exhibited in Fig. S-3B (supplementary

material). The result showed that the maximum extraction efficiency was achieved

when sonication time was increased to 30 s, showing a very fast extraction process.

Therefore, in the following study, the extraction time of 30 s was chosen.

3.3.3 Effect of sample pH and ionic strength


12
The effect of different initial pH values ranging from 2.0 to 12.0 was studied.

The pH values were adjusted by concentrated hydrochloric acid and sodium

hydroxide. From Fig. S-3C (supplementary material), it was found that the extraction

maxima generally occurred in a broad range of pH value from 3.0 to 10.0. Generally,

the pH of the natural solution of nine BPs was close to 5.2. In this study, the sample

solutions were used directly without any pH adjustment.

In order to investigate the influence of the ionic strength, extraction experiments

were carried out in the presence of 0, 0.02, 0.05, 0.1, 0.2, 0.5 and 1.0 mol L-1 NaCl,

respectively. Results (Fig. S-3D, supplementary material) showed that the extraction

efficiency had no significant change with increasing NaCl concentrations (0.02 - 1.0

mol L-1). In this experiment, the sample solutions without adjusting ionic strength

were adopted.

3.3.4 Effect of eluting solvent

Desorption of the target analytes from the sorbents was performed by passing

eluent solution through the filter membrane by a syringe. 5.0 mL of different eluents,

including acetonitrile (ACN), methanol (MeOH), ethyl acetate (CH3COOC2H5),

chloroform (CH3Cl), and acetone (CH3COCH3) were studied. All the eluents

containing analytes were concentrated into 0.5 mL MeOH for HPLC analysis. The

results (Fig. 5A) showed that the highest desorption efficiency for the target analytes

was attained by using MeOH, which may be attributed to the high solubility of BPs in

MeOH. So MeOH was selected as the elution solvent in the subsequent experiments.

3.3.5 Effect of eluent volume

13
The eluent volume affects the sensitivity of the method. During the desorption

process, the effect of eluent volume was studied ranging from 1.0 to 5.0 mL. As could

be seen in Fig. 5B, 4.0 mL of the eluent was enough to desorb the target analytes.

Therefore, 4.0 mL of the eluent was selected for the further experiments.

3.4 Effect of sample volume

The effect of sample volume on quantitative analysis of the analytes was studied

in the range of 10.0-250.0 mL; 10.0 mL sample containing 0.5 g mL-1 of BPs was

diluted to 10.0 to 250.0 mL with deionized water. Then the extraction process was

performed under the optimum conditions. To clearly show the effect of sample

volume on extraction, the values of A/A10 (A: the peak areas at different volumes; A10:

the peak area at volume of 10.0 mL) versus sample volume was plotted (shown in Fig.

S-4, supplementary material). The results showed that the sample in the volumes up to

200.0 mL was quantitatively extracted by the HPCSs, but there was an obvious

decrease in the amount extracted at higher volumes (A/A10 was less than 90%).

Therefore, for the determination of trace quantities of the BPs, a sample volume of

200.0 mL was selected for a high preconcentration factor. The maximum

preconcentration factor could reach 400.

3.5 Inuence of interfering substances

In order to assess the possible analytical applications of the proposed method, the

potential influence of some interferences on the determination of nine BPs in real

samples was examined under the optimal conditions. The results (Table S-1,

supplementary material) suggested that 1000-fold concentration of Na+, K+, NH4+, Cl-,

14
NO3-, SO42-, and 500-fold glucose, 100-fold hydroquinone, resorcinol and 10-fold

PO43- had no obvious influence on the recovery of BPs with deviation below 5%.

3.6 Method evaluation

3.6.1 Analytical performances

Under optimal experimental conditions, a series of experiments were performed

to obtain linear ranges, precision, the limit of detection (LOD) and quantification

(LOQ). All the experiments were performed in triplicate. The working curves were

established for BPS and TDP in the concentration range of 2.0-100.0 g L1; BPF,

BPA, BPAF, BPAP, BPC, TCBPA and TBBPA in the concentration range of 1.0-100.0

g L-1 by analyzing 200.0 mL of spiked deionized water samples. Good linearities

were observed with the determination coefficients (R2) ranging from 0.994 to 0.999.

The limit of detection (LOD=3SD/b, SD is standard deviation of the response, n=9; b:

slope of the calibration curve) and the limit of quantization (LOQ=10SD/b) were

obtained in the range of 0.055-0.529 g L-1 and 0.183-1.763 g L-1, respectively. The

results (Table S-2, supplementary material) showed that the LODs were at low g L-1

level for the target analytes. Additionally, the intra-day precision determined by

analyzing the samples five times in one day was in the range of 1.43.5%. The

inter-day precision achieved by analyzing the samples once a day in three consecutive

days was from 3.1% to 5.2%. The results indicated that the present method had good

repeatability. All values for accuracy and precision were within the recommended

limits.

3.6.2 Application in real sample analysis

15
The optimized SPE method was applied to extract nine BPs from various real

samples including sea water, river water and soft drinks. Fig. 6 presents representative

chromatograms of the sea water sample after pretreatment by SPE with the HPCSs as

sorbents. The average recoveries of the analytes spiked in the samples ranged from

89.6% to 111.5% with relative standard deviations (RSD) less than 6.3% (Table 1).

The results of the experiments confirmed that the method was validated with good

reproducibility, satisfactory precision and accuracy for the determination of BPs from

real samples.

3.7 Comparing with other methods

The performance of the present SPE-HPLC-DAD method based on the prepared

HPCSs was compared with other analytical methods for BPs determination in the

literature. As seen from Table 2, the as-prepared HPCSs sorbent could realize the

simultaneous analysis of nine BPs. The proposed analytical method exhibited lower

LODs than other SPE methods, indicating that the developed method would be a more

appropriate method for the determination of nine BPs at trace levels. In addition,

compared with other sorbents, HPCSs displayed higher extraction efficiency for nine

BPs and lower cost due to lower sorbent dosage used in the present study. Therefore,

the HPCSs were proved to be effective sorbents for simultaneous determination of

nine BPs.

4. CONCLUSIONS

In this work, the synthesis and characterization of monodispersed HPCSs were

achieved prior to their application as extraction sorbent combined with HPLC-DAD

16
for simultaneous determination of nine bisphenols in sea water, river water and soft

drinks. Due to the high specific surface area and good dispersity, the HPCSs exhibited

superior extraction ability for BPs compared to traditional CMs including MWCNTs

and 3D-G. Because of the good results obtained in the present study, this type of CMs

represents an interesting alternative as extraction sorbents for the determination of

organic compounds in different matrixes. Future work might be developed to increase

the applications devoted to the extraction in other samples of interest.

Acknowledgements

This study was supported by the National Nature Science Foundation of China

(NSFC, 51672116), Science and technology foundation of ocean and fisheries of

liaoning province (201408, 201406), General project of scientific research of the

education department of liaoning province (L2015206), Liaoning scientific

instruments service sharing information platform ability construction funds

(201507A003) and the foundation of 211 project for innovative talent training,

liaoning university. The authors also thank their colleagues and other students who

participated in this study.

17
References
[1] D. Guillarme, J. Schappler, S. Rudaz, J.-L. Veuthey, Coupling ultra-high-pressure liquid
chromatography with mass spectrometry, TrAC Trend. Anal. Chem. 29(1) (2010) 15-27.
[2] Z. Zhang, L. Wang, X. Xu, Y. Dong, L. Zhang, Development of a validated HPLC method for the
determination of tenofovir disoproxil fumarate using a green enrichment process, Anal. Methods 7(15)
(2015) 6290-6298.
[3] X. Liu, C. Wang, Q. Wu, Z. Wang, Porous carbon derived from a metalorganic framework as an
efficient adsorbent for the solid-phase extraction of phthalate esters, J. Sep. Sci. 38(22) (2015)
3928-3935.
[4] J. Muoz, M. Gallego, M. Valcrcel, Solid-phase extractiongas chromatographymass
spectrometry using a fullerene sorbent for the determination of inorganic mercury(II),
methylmercury(I) and ethylmercury(I) in surface waters at sub-ng/ml levels, J. Chromatogr. A
1055(12) (2004) 185-190.
[5] L. Wang, Z. Zhang, X. Xu, D. Zhang, F. Wang, L. Zhang, Simultaneous determination of four trace
level endocrine disrupting compounds in environmental samples by solid-phase microextraction
coupled with HPLC, Talanta 142 (2015) 97-103.
[6] L. Wang, M. Zhang, D. Zhang, L. Zhang, New approach for the simultaneous determination
fungicide residues in food samples by using carbon nanofiber packed microcolumn coupled with
HPLC, Food Control 60 (2016) 1-6.
[7] J.M. Jimnez-Soto, S. Crdenas, M. Valcrcel, Evaluation of carbon nanocones/disks as sorbent
material for solid-phase extraction, J. Chromatogr. A 1216(30) (2009) 5626-5633.
[8] Y. Wang, S. Gao, X. Zang, J. Li, J. Ma, Graphene-based solid-phase extraction combined with
flame atomic absorption spectrometry for a sensitive determination of trace amounts of lead in
environmental water and vegetable samples, Anal. Chim. Acta 716 (2012) 112-118.
[9] L. Liu, T. Feng, C. Wang, Q. Wu, Z. Wang, Enrichment of neonicotinoid insecticides from lemon
juice sample with magnetic three-dimensional graphene as the adsorbent followed by determination
with high-performance liquid chromatography, J. Sep. Sci. 37(11) (2014) 1276-1282.
[10] S.-K. Kim, E. Jung, M.D. Goodman, K.S. Schweizer, N. Tatsuda, K. Yano, P.V. Braun,
Self-Assembly of Monodisperse Starburst Carbon Spheres into Hierarchically Organized
Nanostructured Supercapacitor Electrodes, ACS Appl. Mater. Inter. 7(17) (2015) 9128-9133.
[11] C. Saridara, R. Brukh, Z. Iqbal, S. Mitra, Preconcentration of Volatile Organics on
Self-Assembled, Carbon Nanotubes in a Microtrap, Anal. Chem. 77(4) (2005) 1183-1187.
[12] D. Lin, B. Xing, Tannic Acid Adsorption and Its Role for Stabilizing Carbon Nanotube
Suspensions, Environ. Sci. Technol. 42(16) (2008) 5917-5923.
[13] Q. Zhang, J.-Q. Huang, W.-Z. Qian, Y.-Y. Zhang, F. Wei, The Road for Nanomaterials Industry: A
Review of Carbon Nanotube Production, Post-Treatment, and Bulk Applications for Composites and
Energy Storage, Small 9(8) (2013) 1237-1265.
[14] S. Yan, T.-T. Qi, D.-W. Chen, Z. Li, X.-J. Li, S.-Y. Pan, Magnetic solid phase extraction based on
magnetite/reduced graphene oxide nanoparticles for determination of trace isocarbophos residues in
different matrices, J. Chromatogr. A 1347 (2014) 30-38.
[15] A.-H. Lu, G.-P. Hao, Q. Sun, Can Carbon Spheres Be Created through the Stber Method?, Angew.
Chem. Int. Edit. 50(39) (2011) 9023-9025.
[16] J. Liu, N.P. Wickramaratne, S.Z. Qiao, M. Jaroniec, Molecular-based design and emerging
applications of nanoporous carbon spheres, Nat. Mater. 14(8) (2015) 763-774.

18
[17] N.P. Wickramaratne, J. Xu, M. Wang, L. Zhu, L. Dai, M. Jaroniec, Nitrogen Enriched Porous
Carbon Spheres: Attractive Materials for Supercapacitor Electrodes and CO 2 Adsorption, Chem. Mater.
26(9) (2014) 2820-2828.
[18] Y. Zhu, T. Ikoma, N. Hanagata, S. Kaskel, Rattle-Type Fe3O4@SiO2 Hollow Mesoporous Spheres
as Carriers for Drug Delivery, Small 6(3) (2010) 471-478.
[19] C. Merlet, B. Rotenberg, P.A. Madden, P.-L. Taberna, P. Simon, Y. Gogotsi, M. Salanne, On the
molecular origin of supercapacitance in nanoporous carbon electrodes, Nat. Mater. 11(4) (2012)
306-310.
[20] H. Wu, J.-b. Guo, L.-m. Du, H. Tian, C.-x. Hao, Z.-f. Wang, J.-y. Wang, A rapid shaking-based
ionic liquid dispersive liquid phase microextraction for the simultaneous determination of six synthetic
food colourants in soft drinks, sugar- and gelatin-based confectionery by high-performance liquid
chromatography, Food Chem. 141(1) (2013) 182-186.
[21] L.M. Ravelo-Prez, A.V. Herrera-Herrera, J. Hernndez-Borges, M.. Rodrguez-Delgado,
Carbon nanotubes: Solid-phase extraction, J. Chromatogr. A 1217(16) (2010) 2618-2641.
[22] J. Yang, Y. Li, J. Wang, X. Sun, R. Cao, H. Sun, C. Huang, J. Chen, Molecularly imprinted
polymer microspheres prepared by Pickering emulsion polymerization for selective solid-phase
extraction of eight bisphenols from human urine samples, Anal. Chim. Acta 872 (2015) 35-45.
[23] Z. Xu, Z. Yang, Z. Liu, Development of dual-templates molecularly imprinted stir bar sorptive
extraction and its application for the analysis of environmental estrogens in water and plastic samples,
J. Chromatogr. A 1358 (2014) 52-59.
[24] H. Gallart-Ayala, E. Moyano, M.T. Galceran, Recent advances in mass spectrometry analysis of
phenolic endocrine disruptors and related compounds, Mass Spectrom. Rev. 29(5) (2010) 776-805.
[25] M. Fischnaller, R. Bakry, G.K. Bonn, A simple method for the enrichment of bisphenols using
boron nitride, Food Chem. 194 (2016) 149-155.
[26] C. Liao, F. Liu, H. Alomirah, V.D. Loi, M.A. Mohd, H.-B. Moon, H. Nakata, K. Kannan,
Bisphenol S in Urine from the United States and Seven Asian Countries: Occurrence and Human
Exposures, Environ. Sci. Technol. 46(12) (2012) 6860-6866.
[27] X. Yang, Z. Zhang, Y. Fu, Q. Li, Porous hollow carbon spheres decorated with molybdenum
diselenide nanosheets as anodes for highly reversible lithium and sodium storage, Nanoscale 7(22)
(2015) 10198-10203.
[28] W.S. Hummers, R.E. Offeman, Preparation of Graphitic Oxide, J. Am. Chem. Soc. 80(6) (1958)
1339-1339.
[29] C.-X. Gui, Q.-Q. Wang, S.-M. Hao, J. Qu, P.-P. Huang, C.-Y. Cao, W.-G. Song, Z.-Z. Yu,
Sandwichlike Magnesium Silicate/Reduced Graphene Oxide Nanocomposite for Enhanced Pb2+ and
Methylene Blue Adsorption, ACS Appl. Mater. Inter. 6(16) (2014) 14653-14659.
[30] L. Zhang, P. Fang, L. Yang, J. Zhang, X. Wang, Rapid Method for the Separation and Recovery of
Endocrine-Disrupting Compound Bisphenol AP from Wastewater, Langmuir 29(12) (2013) 3968-3975.

19
List of figure captions and table captions
Fig. 1 The scheme of the HPCSs synthesis process

Fig. 2 The TEM image of HPCSs (A); the SEM image of HPCSs (B and C)

Fig. 3 N2 adsorption/desorption isotherms of the HPCSs, 3D-G and MWCNTs (A)

(SBET: BET specific surface area; TPV: Total pore volume; DP: Pore

diameter); FT-IR spectra (B) and EDS spectra (C) of SiO2@C and HPCSs;

Raman spectrum of HPCSs (D)

Fig. 4 The comparison of extraction ability of different CMs: (A) the extraction

efficiency of different CMs; (B) the adsorption capabilitis of different CMs

for 50.0 mg L-1 BPs (volume: 200.0 mL)

Fig. 5 The optimization of desorption condition: (A) the selection of eluent type

(eluent volume: 5.0 mL); (B) the selection of eluent (MeOH) volume

Fig. 6 Representative chromatograms of the sea water extraction solution and sample

spiked with 2.5 g L-1 and 25.0 g L-1 under the determined optimal

conditions

Table 1 Summary of results from analysis of BPs in different samples by HPCSs

based SPE-HPLC method

Table 2 Comparison with other methods

20
Fig. 1

21
Fig. 2

22
Fig. 3

2500 A
Amount adsorbed (cm3 g-1 STP)
SBET TPV Dp
Carbon materials B
(m2 g-1) (cm3 g-1) (nm)
2000 MWCNTs 343.603 0.706 30.202
3D-G 312.889 3.300 3.307 SiO2@C
1500 HPCSs 571.278 1.710 3.494

T%
HPCSs Si-O-Si

1000
MWCNTs
3D-G
500 HPCSs

0
0.0 0.2 0.4 0.6 0.8 1.0 3500 3000 2500 2000 1500 1000 500
Relative pressure (P/P0) Wavenumber(cm-1)
2500
a Si
C D
D-band G-band
2000

O
1500
Intensity (a.u.)

C
1000
b
500

-500
0 100 200 300 400 1000 1500 2000 2500
Energy (eV) Raman shift (cm-1)

23
Fig. 4
MWCNTs MWCNTs
3D-G 3D-G
HPCSs HPCSs
200
100 A B

160
Extraction efficiency(%)

80

120

qe(mg g-1)
60

40 80

20 40

0 0
BPS BPF TDP BPA BPAF BPAP BPC TCBPA TBBPA BPS BPF TDP BPA BPAF BPAP BPC TCBPA TBBPA

24
Fig. 5

700 A BPS
700 B BPS
BPF BPF
TDP TDP
600 600
BPA BPA
BPAF BPAF
500 500

Peak area
BPAP
Peak area

BPAP
BPC BPC
400 400
TCBPA TCBPA
TBBPA 300 TBBPA
300

200 200

100 100

0 0
MeOH ACN CH3COOC2H5 CH3Cl CH3COCH3 1 2 3 4 5
Eluent Volume of eluent (mL)

25
Fig. 6

Spiked with 25 g L-1 TCBPA


Spiked with 2.5 g L-1 TBBPA
BPAP
Blank
BPAF
Peak height (mAU)

BPC

BPA

Peak height (mAU)


BPF
TDP
BPS

4 6 8 10
4 6 8 10 12 14 16 18 20 Retention time (min)
Retention time (min)

26
Table 1 Summary of results from analysis of BPs in different samples by HPCSs

based SPE-HPLC method

sea water river water grape juice


Added
Analytes Found Recovery RSD Found Recovery RSD Found Recovery RSD
(g L-1)
(g L-1) (%) (%) (g L-1) (%) (%) (g L-1) (%) (%)
BPS 0 nd - - nd - - nd - -
2.5 2.47 98.8 2.0 2.55 102.0 2.7 2.61 104.4 2.3
25.0 25.90 103.6 3.7 23.92 95.7 5.2 27.63 110.5 5.6
BPF 0 nd - - nd - - nd - -
2.5 2.40 96.0 3.3 2.42 96.8 4.5 2.28 91.2 4.4
25.0 23.80 95.2 1.6 26.13 104.5 5.1 26.33 105.3 4.0
TDP 0 nd - - nd - - nd - -
2.5 2.55 102.0 3.1 2.63 105.2 0.8 2.36 94.4 5.5
25.0 24.10 96.4 3.5 27.03 108.1 3.9 23.72 94.9 3.2
BPA 0 nd - - nd - - nd - -
2.5 2.66 106.4 4.9 2.72 108.8 2.9 2.41 96.4 3.7
25.0 24.60 98.4 4.7 23.99 96.0 4.4 26.29 105.2 2.8
BPAF 0 nd - - nd - - nd - -
2.5 2.71 108.4 3.7 2.44 97.6 5.3 2.70 108.0 4.4
25.0 23.10 92.4 5.6 23.83 95.3 4.9 25.03 100.1 3.2
BPAP 0 nd - - nd - - nd - -
2.5 2.54 101.6 4.3 2.60 104.0 5.0 2.24 89.6 1.8
25.0 25.10 100.4 4.1 25.09 100.4 3.9 23.28 93.1 4.3
BPC 0 nd - nd - - nd - -
2.5 2.31 92.4 5.6 2.53 101.2 4.0 2.67 106.8 2.6
25.0 24.62 98.5 3.8 23.22 92.9 5.7 26.38 105.5 2.7
TCBPA 0 nd - - nd - - nd - -
2.5 2.69 107.6 2.6 2.37 94.8 6.3 2.29 91.6 2.6
25.0 27.88 111.5 4.6 23.07 92.3 3.3 27.04 108.2 3.3
TBBPA 0 nd - - nd - - nd - -
2.5 2.29 91.6 3.9 2.69 107.6 2.2 2.32 92.8 3.4
25.0 25.38 101.5 2.1 24.27 97.1 2.4 23.73 94.9 4.6

27
Table 2 Comparison with other methods
Sorbent dosage Target analyte LODs
Analytical method Sorbent a -1
Refs
(g L ) number (g L-1)
SPE-HPLC-DAD MIPMS 30.0 8 1.2-2.2 22
SPME-HPLC-DAD MWCNTs 0.16 4 0.10-0.30 5
SPE-HPLC-FLD BN 0.5-2.0 5 0.23-0.28 25
SPE-HPLC-DAD HPCSs 0.025 9 0.05-0.53 This work

a
MIPMS, molecularly imprinted polymer microsphere; BN, hexagonal boron nitride

Highlights:

The monodispersed hollow porous carbon spheres - HPCSs were controlled

synthesized.

The synthesized HPCSs exhibited superior extraction ability to bisphenols

compared with MWCNTs and 3D-G.

HPCSs were successfully applied as solid-phase extraction adsorbent for the

enrichment and simultaneous determination of nine biphenols in real samples.


28

Potrebbero piacerti anche