Next Article in Journal
Elaboration of Highly Modified Stainless Steel/Lead Dioxide Anodes for Enhanced Electrochemical Degradation of Ampicillin in Water
Next Article in Special Issue
Development and Application of Green or Sustainable Strategies in Analytical Chemistry
Previous Article in Journal
Understanding the Effects of Inlet Structure on Separation Performance Based on Axial Velocity Wave Zone Characteristics
Previous Article in Special Issue
Evaluation of Pulsed Electric Field-Assisted Extraction on the Microstructure and Recovery of Nutrients and Bioactive Compounds from Mushroom (Agaricus bisporus)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Preparation of Phenyboronic-Acid-Functionalized Fe3O4 Magnetic Nanoparticles for the Selective Adsorption of Ortho-Dihydroxy-Containing Compounds

Department of Chemistry, Yunnan Normal University, Kunming 650500, China
*
Authors to whom correspondence should be addressed.
Separations 2023, 10(1), 4; https://doi.org/10.3390/separations10010004
Submission received: 29 October 2022 / Revised: 27 November 2022 / Accepted: 13 December 2022 / Published: 21 December 2022

Abstract

:
A new facile strategy was designed to prepare the phenyboronic acid-functionalized Fe3O4 magnetic nanoparticles (Fe3O4@PBA) via direct silanization and thiol-ene click chemistry for the selective adsorption of ortho-dihydroxy-containing compounds. The three kinds of Fe3O4@PBA nanoparticles obtained showed excellent adsorption capacity and selectivity for ortho-dihydroxy-containing compounds including adenosine and o-dihydroxybenzene. Among them, the Fe3O4@MPS@MPBA exhibited the highest adsorption capacity and selectivity for adenosine and o-dihydroxybenzene, followed by Fe3O4@MPTES@AAPBA and Fe3O4@MPTES@VPBA. A synthesis method of superparamagnetic and boronate affinity nanocomposites with mild reaction conditions and simple process has been developed, which also provides a novel way for the synthesis of other types of enrichment materials of ortho-dihydroxy-containing compounds.

Graphical Abstract

1. Introduction

Due to the advantages of superparamagnetism, stability and low cytotoxicity [1,2,3], magnetic nanomaterials are widely used in various fields, such as pollutant detection and separation [4,5,6,7], drug delivery [8,9,10], magnetic resonance imaging [11,12,13] and biosensing [14,15,16]. Such a wide range of applications is attributed to the various functionalizations of magnetic nanomaterials, and phenylboric acid is one of the important functional modification groups. Phenylboric acid can selectively adsorb ortho-dihydroxy-containing compounds through the formation of reversible five- or six-membered cycloesters. The adsorption and desorption can be conveniently controlled by adjusting the pH value, which gives boric acid affinity materials incomparable advantages in the recognition and separation of biomolecules [17,18,19,20,21,22,23,24,25]. Accordingly, it has become a challenge to develop a facile modification strategy for the preparation of magnetic nanomaterials modified by phenylboric acid. Click chemistry has become a powerful means to solve the above challenge because of its high reaction efficiency, high reliability and good selectivity [26,27,28,29].
In this study, the direct silylation and thiol-ene click reaction were used for functionalized modification of Fe3O4 magnetic nanoparticles. The adsorption capacity and selectivity of the three kinds of prepared phenyboronic-acid-functionalized Fe3O4 magnetic nanoparticles (Fe3O4@PBA) toward ortho-dihydroxy-containing compounds were evaluated by adsorption experiments of adenosine and o-dihydroxybenzene.

2. Materials and Methods

2.1. Reagents and Material

Ferric trichloride hexahydrate (FeC13·6H2O), ammonium hydrogen carbonate, sodium chloride, ethylene glycol, anhydrous sodium acetate, glacial acetic acid, methanol, ethanol and aqueous ammonia (25 wt%) were obtained from FengChuan Fine Chemical Research Institute (Tianjin, China). Adenosine, deoxyadenosine, azodiisobutyronitrile (AIBN), 3-mercaptopropyltrimethoxysilane (MPTES) and 4-vinylphenylboronic acid (VPBA) were obtained from Adamas-beta (Basel, Switzerland). 3-methacryloxypropyltrimethoxysilane (MPS) and 3-acrylamidophenylboronic acid (AAPBA) were purchased from Sigma-Aldrich (St. Louis, MO, USA). 4-Mercaptophenylboronic acid (MPBA), o-dihydroxybenzene, m-dihydroxybenzene, and p-dihydroxybenzene were obtained from Aladdin (Shanghai, China). Deionised water was used to prepare all buffer and analyte solutions.

2.2. Synthesis of Fe3O4@PBA Nanoparticles

The general scheme for the synthesis of magnetic nanoparticles modified by phenylboronic acid (Fe3O4@PBA) is illustrated in Figure 1. The Fe3O4 nanoparticles was synthesized by a previously reported solvothermal method [30]. The Fe3O4 nanoparticles were functionalized with the sulfydryl and vinyl groups through the direct silanization using MPTES and MPS [31]. Subsequently, the functionalization of the nanoparticles based on phenylboric acid was achieved by the thiol-ene click reaction using 4-vinylphenylboronic acid (VPBA), 3-acrylamidophenylboronic acid (AAPBA) and 4-mercaptophenylboronic acid (MPBA) [32].

2.2.1. Synthesis of Fe3O4 Nanoparticles

A solution containing anhydrous sodium acetate (7.2 g), polyethylene glycol (2.0 g), FeCl3·6H2O (2.7 g) and ethylene glycol (80.0 mL) was stirred at 50 °C for 15 min. Then the solution was transferred to a PTFE reactor and heated at 200 °C for 8 h. The products were collected with magnets and washed repeatedly using deionized water and ethanol, respectively. Finally, the products were dried overnight in a vacuum at 50 °C.

2.2.2. Synthesis of Fe3O4@MPTES

The Fe3O4 nanoparticles (200.0 mg), ethanol (25.0 mL) and deionized water (20.0 mL) were added into a 250 mL two-neck round-bottom flask, and the mixture was ultrasonically dispersed for 30 min. Then, 1.0 mL of glacial acetic acid was slowly added under mechanical agitation (350 rpm) to adjust the pH value of the solution to about 5, and stirring continued for 30 min. Subsequently, the direct silanization reaction was initiated by the dropwise addition of 5.0 mL MPTES, which was maintained in a water bath at 50 °C for 24 h. The products were separated with magnets, and washed with deionized water and ethanol for several times. Then, the products were dried overnight at 50 °C to obtain Fe3O4@MPTES nanoparticles.

2.2.3. Synthesis of Fe3O4@MPS

The Fe3O4 nanoparticles (200.0 mg), ethanol (40.0 mL), deionized water (10.0 mL) and ammonium hydroxide (1.5 mL) were added into a 250 mL two-neck round-bottom flask, and the mixture was ultrasonically dispersed for 30 min. Whereafter, 500 μL of MPS was dropwise added under mechanical agitation (350 rpm) to initiate the direct silanization reaction. After reacting at 50 °C for 24 h, the obtained Fe3O4@MPS was repeatedly washed using deionized water and ethanol and dried overnight at 50 °C.

2.2.4. Phenyboronic-Acid-Functionalized Fe3O4@MPS and Fe3O4@MPTES

200.0 mg of Fe3O4@MPS or Fe3O4@MPTES was ultrasonically dispersed in a solution containing 50 mL of ethanol and 200.0 mg of MPBA (400.0 mg of AAPBA or 200.0 mg of VPBA for Fe3O4@MPTES) for 30 min. Then the thiol-ene click reaction was initiated by the addition of AIBN (20.0) mg under nitrogen protection, mechanical stirring (400 rpm) and 50 °C for 24 h. The products were separated with magnets, and washed using deionized water and ethanol for several times. The vacuum drying process was maintained at 50 °C for 12 h. Finally, the obtained products were named as Fe3O4@MPTES@VPBA, Fe3O4@MPTES@AAPBA and Fe3O4@MPS@MPBA, respectively.

2.3. Binding Experiments

A static adsorption method was used to evaluate the adsorption capacity of the three kinds of Fe3O4@PBA nanoparticles: Briefly, Fe3O4@PBA (2.0 mg) was dispersed by ultrasound in test solution (0.2 mL) with different concentrations (0.1~1.0 mg/mL) which was prepared with an ammonium bicarbonate (50 mM, pH = 8.5, containing 500 mM NaCl) as buffer solution, and shaken at room temperature (1000 rpm) for 20~140 min. The concentration of the adsorbate in the test solution after adsorption was measured by UV–Vis spectrophotometry. Then the adsorption capacity (Qe) was calculated according to the equation below:
Q e = V ( C 0 C e ) m × 1000
where C0 (mg/mL) is the initial concentration of the adsorbate in the test solution; Ce (mg/mL) is the equilibrium concentration of the adsorbate; V (mL) is the volume of the adsorbate solution; m (mg) is the mass of Fe3O4@PBA. Contrast adsorption experiments were also performed with Fe3O4@MPTES and Fe3O4@MPS under the same conditions.
The Scatchard equation was employed to investigate the binding properties of the Fe3O4@PBA nanoparticles:
Q C = ( Q m a x Q ) K D
where Q (mg/g) is the equilibrium adsorption capacity of the material to the adsorbate, C (mg/mL) is the adsorbate concentration remaining in the test solution after adsorption equilibrium, Qmax (mg/g) is the maximum apparent binding amount, KD is the equilibrium dissociation constant.

2.4. Selectivity Experiments

The selective adsorption capacity of Fe3O4@PBA nanoparticles can be investigated by adsorption experiments on mixed solutions of two or more adsorbates with similar structure but different in the presence of cis-diol. Two mixed solutions (1.0 mg/mL) with equal mass ratios of adsorbates were used in the selective adsorption experiments of Fe3O4@PBA nanoparticles, including adenosine and deoxyadenosine, and three dihydroxybenzene isomers. The procedure of adsorption experiments is the same as in Section 2.3, except that the Fe3O4@PBA nanoparticles were eluted by acetic acid buffer solution (0.05 mL, pH = 3.0) for 60 min after adsorption and rinsed. The eluent was analysed by HPLC. The Chromatographic analysis was carried out using a P230II HPLC with a Shim-pack C18 column (150 mm × 4.6 mm, 5 μm), and the mobile phase was methanol-water (12:88, v/v) at the flow rate of 1.0 mL/min. The column temperature was maintained at 30 °C, and the wavelength of the UV detector was 260 nm for adenosine and deoxyadenosine (280 nm for dihydroxybenzene isomers).

3. Results and Discussion

3.1. Characterization of Fe3O4@PBA Nanoparticles

Figure 2 is the transmission electron microscope (TEM) images of Fe3O4@PBA nanoparticles. As shown in Figure 2, three kinds of nanoparticles formed a core-shell structure including a core composed of Fe3O4 and a modified layer about 5 nm thick.
The results of the scanning transmission electron microscopy (STEM) analysis of element distribution (Figure 3) shows that B, Si, N and S are uniformly distributed around the core composed of Fe and O, indicating that the modified layer composed of organosilane and phenylboric acid has been successfully introduced on the particle surface.
The above elements’ distribution of Fe3O4@PBA nanoparticles was also confirmed by X-ray photoelectron spectroscopy (XPS). As shown in Figure 4, strong signal peaks at 530 eV and 284.8 eV indicate that there are abundant oxygen-containing groups and carbon elements on the surface of the nanoparticles, and the peaks at 710.5 eV, 399.6 eV, 285.0 eV, 191.4 eV, 163.2 eV, and 102.1 eV correspond to Fe2p, N1s, C1s, B1s, S2p3 and Si2p2, respectively. The signal peaks of S in Figure 4a,b shows that the hydrosulphonyl groups were introduced to the surface of Fe3O4 nanoparticles by the coating of MPTES, and the signal peak of N in Figure 4b and the signal peak of B in Figure 4a–c all indicate that phenylboric acid has been successfully modified on the particle surface.
The weight loss of different modified nanoparticles can provide more supporting evidence for the modification process under nitrogen atmosphere with a heating rate of 10 °C/min. As can be seen from Figure 5, the Fe3O4 nanoparticles had no significant weight loss during the heating process. However, with the modification of organosilane and the introduction of phenylboronic acid groups, the weight loss rate of the modified nanoparticles increased.
The room temperature magnetism of the prepared Fe3O4@PBA nanoparticles was analysed by using the vibration sample magnetometer (VSM). As can be seen from Figure 6, the room temperature saturation magnetic curves of the three kinds of Fe3O4@PBA nanoparticles have no obvious hysteresis loop and coercivity. In addition, nanoparticles dispersed in water can be reaggregated within 15 s under an applied magnetic field. The above analysis results and experimental phenomena shown that the prepared Fe3O4@PBA nanoparticles have excellent superparamagnetism.

3.2. Binding Properties of Fe3O4@PBA Nanoparticles

3.2.1. Thermodynamics of Adsorption

The adsorption isotherms of the Fe3O4@MPTES@VPBA toward different adsorbates are shown in Figure 7. As can be seen from the adsorption isotherms, when the concentration of adsorbate is low, the adsorption capacity of the Fe3O4@MPTES@VPBA on the adsorbate increases synchronously with the concentration of the adsorbate. However, due to the high concentration of adsorbate, the boric acid binding site on the surface of the material tends to saturate, which makes the increasing trend of adsorption capacity slow down gradually. When the initial concentration of adsorbed substance was 1.0 mg/mL, the adsorption capacity basically reaches the maximum. Similarly, the isothermal adsorption curves of Fe3O4@MPTES@AAPBA and Fe3O4@MPS@MPBA shown the same trend of adsorption capacity as that of Fe3O4@MPTES@VPBA.
The adsorption isotherm of Fe3O4@MPS, Fe3O4@MPTES, and three kinds of Fe3O4@PBA nanoparticles for the adsorption of adenosine and o-dihydroxybenzene were analysed by Scatchard equation [33,34]. The fitting curves of Fe3O4@MPTES@VPBA and Fe3O4@MPTES are shown in Figure 8. As can be seen from Figure 8, due to the introduction of phenylboric acid high-affinity sites, the fitting curves of Fe3O4@MPTES@VPBA are composed of two straight lines representing high-affinity sites and low-affinity sites, respectively (Figure 8a,c). In contrast, the fitting curves of Fe3O4@MPTES show a linear relationship over the whole concentration range due to the absence of high-affinity sites (Figure 8b,d). Actually, the three kinds of Fe3O4@PBA nanoparticles have similar fitting analysis results. According to Table 1, there are two binding sites of high affinity and low affinity on the three kinds of Fe3O4@PBA nanoparticles. The equilibrium dissociation constant (Kd) of the high-affinity binding sites for the adsorption of adenosine and o-dihydroxybenzene was significantly lower than those of the low-affinity sites, indicating that the specific adsorption caused by boric acid affinity had a higher affinity than the non-specific adsorption caused by hydrogen bonding and electrostatic interaction.
Boric acid groups can specifically form ester rings with ortho-dihydroxy-containing compounds, resulting in selective adsorption. Unlike adenosine, deoxyadenosine does not have adjacent hydroxyl groups. The adsorption of deoxyadenosine by Fe3O4@PBA depends only on hydrogen bonding and electrostatic interaction, which results in the adsorption capacity of deoxyadenosine being significantly lower than that of adenosine.
Of the three isomers of dihydroxybenzene, o-Dihydroxybenzene is the only one that has adjacent hydroxyl groups, and it forms ester rings with boric acid much more easily than m-dihydroxybenzene and p-dihydroxybenzene. Therefore, the adsorption capacity and selectivity of Fe3O4@PBA for o-dihydroxybenzene are greater than that of m-dihydroxybenzene and p-dihydroxybenzene.

3.2.2. Kinetics of Adsorption

As can be seen from the adsorption kinetics curve of Fe3O4@MPTES@VPBA (Figure 9), at the initial stage of adsorption, a relatively stable complex was formed between the adsorbate and the boric acid binding site, which causes the adsorption capacity to rise rapidly. When most of the boric acid binding sites on the surface are occupied, it becomes difficult for the adsorbate to bind to the deeper binding sites, resulting in a slow increase in the amount of adsorption. With the continuous progress of the adsorption, all binding sites in the material are saturated, and the adsorption capacity tends to be stable. As shown in Figure 9, all the three kinds of prepared Fe3O4@PBA nanoparticles showed the same trend of adsorption capacity and reached their maximum adsorption capacity at 80~100 min.

3.3. Selectivity

A solution of adenosine or deoxyadenosine (1.0 mg/L) was adsorbed by the three prepared Fe3O4@PBA nanoparticles, respectively. The adsorption selectivity factor (α = Qadenosine/Qdeoxyadenosine) was calculated. Similarly, the adsorption selectivity factors of dihydroxybenzene isomers were also obtained by the same method. The calculation results are shown in Table 2.
According to the Table 2, the adsorption capacity of the three kinds of Fe3O4@PBA nanoparticles toward adenosine was significantly higher than that of deoxyadenosine, and the adsorption selectivity factor reached more than 4.5. Furthermore, the adsorption selectivity factor for o-dihydroxybenzene ranged from 2.18 to 3.17, indicating that all three kinds of Fe3O4@PBA nanoparticles offered excellent adsorption selectivity toward adenosine and o-dihydroxybenzene. Particularly, the Fe3O4@MPS@MPBA exhibited the highest adsorption capacity and selectivity factor for adenosine and o-dihydroxybenzene.
The adsorption selectivity was further confirmed by HPLC analysis of the eluents after adsorption of the mixed solutions. Figure 10 presents the HPLC chromatograms of the eluents of the mixed solution (1.0 mg/mL) with equal mass ratio of adenosine and deoxyadenosine adsorbed by three kinds of Fe3O4@PBA nanoparticles, respectively. As can be seen from Figure 10, the chromatographic peak area of adenosine in the eluents was significantly higher than that of deoxyadenosine, which proved that three kinds of Fe3O4@PBA nanoparticles also had obvious adsorption selectivity for adenosine in the mixed solution of adenosine and deoxyadenosine.
The HPLC chromatograms of the eluents of the mixed solution (1.0 mg/mL) with an equal mass ratio of o-dihydroxybenzene, m-dihydroxybenzene and p-dihydroxybenzene adsorbed by the three kinds of Fe3O4@PBA nanoparticles, respectively, are shown in Figure 11. Obviously, the peak area of o-dihydroxybenzene was the largest in all the eluents. It was also significantly different from that of m-dihydroxybenzene and p-dihydroxybenzene, further indicating that the three kinds of Fe3O4@PBA nanoparticles also showed the adsorption selectivity of o-dihydroxybenzene in the mixed solution of the dihydroxybenzene isomers.

4. Conclusions

In this study, a new modification strategy for Fe3O4 magnetic nanoparticles with phenylboric acid was designed using the direct silanizing method and thiol-ene click reaction. Based on this new strategy, three kinds of phenylboric-acid-modified magnetic nanoparticles (Fe3O4@PBA) were prepared by a simple process under mild reaction conditions. The three kinds of Fe3O4@PBA nanoparticles obtained showed excellent adsorption capacity and selectivity for ortho-dihydroxy-containing compounds, including adenosine and o-dihydroxybenzene. The reactivity of acrylates was higher than that of acrylamide and styrene among the double-bond monomers that produce thiol-ene click chemistry, which resulted in a relatively larger number of phenylboronic acid binding sites on the surface of magnetic nanoparticles modified with 4-mercaptophenylboronic acid (MPBA). Therefore, Fe3O4@MPS@MPBA exhibited the highest adsorption capacity and selectivity factor for adenosine and o-diphenol among the three kinds of Fe3O4@PBA.
Furthermore, boric acid with low pka value should be used in the modification of nanoparticles to reduce the bonding pH value of Fe3O4@PBA, which enables Fe3O4@PBA to selectively adsorb glycopeptides and glycoproteins in biological samples.

Author Contributions

Conceptualization, L.Y.; methodology, S.X.; validation, A.D.; investigation, H.Z.; data curation, J.Z.; writing—original draft preparation, H.Z.; writing—review and editing, B.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Applied Basic Research Foundation of Yunnan Province, grant number 202201AT070029.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study.

References

  1. Prucek, R.; Tucek, J.; Kilianova, M.; Panacek, A.; Kvitek, L.; Filip, J.; Kolar, M.; Tomankova, K.; Zboril, R. The targeted antibacterial and antifungal properties of magnetic nanocomposite of iron oxide and silver nanoparticles. Biomaterials 2011, 32, 4704–4713. [Google Scholar] [CrossRef] [PubMed]
  2. Erathodiyil, N.; Ying, J.Y. Functionalization of inorganic nanoparticles for bioimaging applications. Acc. Chem. Res. 2011, 44, 925–935. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, G.; Gao, J.H.; Ai, H.; Chen, X.Y. Applications and potential toxicity of magnetic iron oxide nanoparticles. Small 2013, 9, 1533–1545. [Google Scholar] [CrossRef] [PubMed]
  4. Babamiri, B.; Salimi, A.; Hallaj, R. Switchable electrochemiluminescence aptasensor coupled with resonance energy transfer for selective attomolar detection of Hg2+ via CdTe@CdS/dendrimer probe and Au nanoparticle quencher. Biosens. Bioelectron. 2018, 102, 328–335. [Google Scholar] [CrossRef] [PubMed]
  5. Moazzen, M.; Ahmadkhaniha, R.; Gorji, M.E.; Yunesian, M.; Rastkari, N. Magnetic solid-phase extraction based on magnetic multi-walled carbon nanotubes for the determination of polycyclic aromatic hydrocarbons in grilled meat samples. Talanta 2013, 115, 957–965. [Google Scholar] [CrossRef] [PubMed]
  6. Facchi, D.P.; Cazetta, A.L.; Canesin, E.A.; Almeida, V.C.; Bonafé, E.G.; Kipper, M.J.; Martins, A.F. New magnetic chitosan/alginate/Fe3O4@SiO2 hydrogel composites applied for removal of Pb(II) ions from aqueous systems. Chem. Eng. J. 2018, 337, 595–608. [Google Scholar] [CrossRef]
  7. Fan, R.; Min, H.; Hong, X.; Yi, Q.; Liu, W.; Zhang, Q.; Luo, Z. Plant tannin immobilized Fe3O4@SiO2 microspheres: A novel and green magnetic bio-sorbent with superior adsorption capacities for gold and palladium. J. Hazard. Mater. 2019, 364, 780–790. [Google Scholar] [CrossRef]
  8. Cai, Y.R.; Pan, H.H.; Xu, X.R.; Hu, Q.H.; Li, L.; Tang, R.K. Ultrasonic controlled morphology transformation of hollow calcium phosphate nanospheres: A smart and biocompatible drug release system. Chem. Mater. 2007, 19, 3081–3083. [Google Scholar] [CrossRef]
  9. Dobson, J. Gene therapy progress and prospects: Magnetic nanoparticle-based gene delivery. Gene Ther. 2006, 13, 283–287. [Google Scholar] [CrossRef] [Green Version]
  10. Zhang, Z.; Niu, N.; Gao, X.; Han, F.; Chen, Z.; Li, S.; Li, J. A new drug carrier with oxygen generation function for modulating tumor hypoxia microenvironment in cancer chemotherapy. Colloids Surf. B. Biointerfaces 2019, 173, 335–345. [Google Scholar] [CrossRef] [PubMed]
  11. Justin, R.; Tao, K.; Román, S.; Chen, D.; Xu, Y.; Geng, X.; Ross, I.M.; Grant, R.T.; Pearson, A.; Zhou, G.; et al. Photoluminescent and superparamagnetic reduced graphene oxide-iron oxide quantum dots for dual-modality imaging, drug delivery and photothermal therapy. Carbon 2016, 97, 54–70. [Google Scholar] [CrossRef] [Green Version]
  12. Sonvico, F.; Dubernet, C.; Colombo, P.; Couvreur, P. Metallic colloid nanotechnology, applications in diagnosis and therapeutics. Curr. Pharm. Des. 2005, 11, 2095–2105. [Google Scholar] [CrossRef] [PubMed]
  13. Xu, C.; Sun, S. Monodisperse magnetic nanoparticles for biomedical applications. Polym. Int. 2007, 56, 821–826. [Google Scholar] [CrossRef]
  14. Villalonga, M.L.; Borisova, B.; Arenas, C.B.; Villalonga, A.; Arévalo-Villena, M.; Sánchez, A.; Pingarrón, J.M.; Briones-Pérez, A.; Villalonga, R. Disposable electrochemical biosensors for Brettanomyces bruxellensis and total yeast content in wine based on core-shell magnetic nanoparticles. Sensor. Actuat. B Chem. 2019, 279, 15–21. [Google Scholar] [CrossRef]
  15. Gu, H.; Ho, P.L.; Tsang, K.W.; Wang, L.; Xu, B. Using biofunctional magnetic nanoparticles to capture vancomycin-resistant enterococci and other gram-positive bacteria at ultralow concentration. J. Am. Chem. Soc. 2003, 125, 15702–15703. [Google Scholar] [CrossRef] [PubMed]
  16. Jiang, W.; Wu, L.; Duan, J.; Yin, H.; Ai, S. Ultrasensitive electrochemiluminescence immunosensor for 5-hydroxymethylcytosine detection based on Fe3O4@SiO2 nanoparticles and PAMAM dendrimers. Biosens. Bioelectron. 2018, 99, 660–666. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, Z.; He, H. Synthesis and applications of boronate affinity materials: From class selectivity to biomimetic specificity. Acc. Chem. Res. 2017, 50, 2185–2193. [Google Scholar] [CrossRef] [Green Version]
  18. Brooks, W.L.; Sumerlin, B.S. Synthesis and applications of boronic acid-containing polymers: From materials to medicine. Chem. Rev. 2016, 116, 1375–1397. [Google Scholar] [CrossRef]
  19. Li, D.J.; Chen, Y.; Liu, Z. Boronate affinity materials for separation and molecular recognition: Structure, properties and applications. Chem. Soc. Rev. 2015, 44, 8097–8123. [Google Scholar] [CrossRef]
  20. Nishiyabu, R.; Kubo, Y.; James, T.D.; Fossey, J.S. Boronic acid building blocks: Tools for sensing and separation. Chem. Commun. 2011, 47, 1106–1123. [Google Scholar] [CrossRef]
  21. Xing, R.; Wang, S.; Bie, Z.; He, H.; Liu, Z. Preparation of molecularly imprinted polymers specific to glycoproteins, glycans and monosaccharides via boronate affinity controllable-oriented surface imprinting. Nat. Protoc. 2017, 12, 964–987. [Google Scholar] [CrossRef] [PubMed]
  22. Luo, J.; Huang, J.; Cong, J.J.; Wei, W.; Liu, X.Y. Double recognition and selective extraction of glycoprotein based on the molecular imprinted graphene oxide and boronate affinity. ACS Appl. Mater. Interfaces 2017, 9, 7735–7744. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, W.; Liu, W.; Li, P.; Xiao, H.B.; Wang, H.; Tang, B. A fluorescence nanosensor for glycoproteins with activity based on the molecularly imprinted spatial structure of the target and boronate affinity. Angew. Chem. Int. Ed. Engl. 2014, 53, 12489–12493. [Google Scholar] [CrossRef] [PubMed]
  24. Wei, J.R.; Ni, Y.L.; Zhang, W.; Zhang, Z.Q.; Zhang, J. Detection of glycoprotein through fluorescent boronic acid-based molecularly imprinted polymer. Anal. Chim. Acta 2017, 960, 110–116. [Google Scholar] [CrossRef]
  25. Ye, J.; Chen, Y.; Liu, Z. A boronate affinity sandwich assay: An appealing alternative to immunoassays for the determination of glycoproteins. Angew. Chem. Int. Ed. Engl. 2014, 53, 10386–10389. [Google Scholar] [CrossRef]
  26. Hoyle, C.E.; Bowman, C.N. Thiol-ene click chemistry. Angew. Chem. Int. Ed. Engl. 2010, 49, 1540–1573. [Google Scholar] [CrossRef]
  27. Such, G.K.; Johnston, A.P.R.; Liang, K.; Caruso, F. Synthesis and functionalization of nanoengineered materials using click chemistry. Prog. Polym. Sci. 2012, 37, 985–1003. [Google Scholar] [CrossRef]
  28. Hayase, G.; Kanamori, K.; Hasegawa, G.; Maeno, A.; Kaji, H.; Nakanishi, K. A superamphiphobic macroporous silicone monolith with marshmallow-like flexibility. Angew. Chem. Int. Ed. Engl. 2013, 52, 10788–10791. [Google Scholar] [CrossRef]
  29. Li, P.; Wang, X.; Zhao, Y. Click chemistry as a versatile reaction for construction and modification of metal-organic frameworks. Coord. Chem. Rev. 2019, 380, 484–518. [Google Scholar] [CrossRef]
  30. Deng, H.; Lin, L.X.; Qing, P.; Wang, X.; Chen, J.P.; Li, Y.D. Monodisperse magnetic single-crystal ferrite microspheres. Angew. Chem. Int. Ed. 2005, 44, 2782–2785. [Google Scholar] [CrossRef]
  31. Bi, C.; Zhang, S.; Li, Y.; He, X.; Chen, L.; Zhang, Y. Boronic acid-functionalized iron oxide magnetic nanoparticles via distillation-precipitation polymerization and thiol-yne click chemistry for the enrichment of glycoproteins. New J. Chem. 2018, 42, 17331–17338. [Google Scholar] [CrossRef]
  32. Zhang, S.; He, X.; Chen, L.; Zhang, Y. Boronic acid functionalized magnetic nanoparticles via thiol–ene click chemistry for selective enrichment of glycoproteins. New J. Chem. 2014, 38, 4212. [Google Scholar] [CrossRef]
  33. Zhu, S.; Xia, M.; Chu, Y.; Khan, M.A.; Lei, W.; Wang, F.; Muhmood, T.; Wang, A. Adsorption and Desorption of Pb(II) on l-Lysine Modified Montmorillonite and the simulation of Interlayer Structure. Appl. Clay Sci. 2019, 169, 40–47. [Google Scholar] [CrossRef]
  34. Zhu, S.; Khan, M.A.; Kameda, T.; Xu, H.; Wang, F.; Xia, M.; Yoshioka, T. New insights into the capture performance and mechanism of hazardous metals Cr(3+) and Cd(2+) onto an effective layered double hydroxide based material. J. Hazard Mater. 2022, 426, 128062. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of synthesis of phenylboronic-acid-modified magnetic nanoparticles.
Figure 1. Schematic illustration of synthesis of phenylboronic-acid-modified magnetic nanoparticles.
Separations 10 00004 g001
Figure 2. TEM images of the phenylboronic-acid-modified magnetic nanoparticles: Fe3O4@MPTES@VPBA (a,d), Fe3O4@MPTES@AAPBA (b,e), Fe3O4@MPS@MPBA (c,f).
Figure 2. TEM images of the phenylboronic-acid-modified magnetic nanoparticles: Fe3O4@MPTES@VPBA (a,d), Fe3O4@MPTES@AAPBA (b,e), Fe3O4@MPS@MPBA (c,f).
Separations 10 00004 g002
Figure 3. STEM elemental mapping images of the phenylboronic-acid-modified magnetic nanoparticles Fe3O4@MPTES@VPBA (a), Fe3O4@MPTES@AAPBA (b) and Fe3O4@MPS@MPBA (c).
Figure 3. STEM elemental mapping images of the phenylboronic-acid-modified magnetic nanoparticles Fe3O4@MPTES@VPBA (a), Fe3O4@MPTES@AAPBA (b) and Fe3O4@MPS@MPBA (c).
Separations 10 00004 g003
Figure 4. XPS spectra of the phenylboronic-acid-modified magnetic nanoparticles (a) Fe3O4@MPTES@VPBA, (b) Fe3O4@MPTES@AAPBA and (c) Fe3O4@MPS@MPBA.
Figure 4. XPS spectra of the phenylboronic-acid-modified magnetic nanoparticles (a) Fe3O4@MPTES@VPBA, (b) Fe3O4@MPTES@AAPBA and (c) Fe3O4@MPS@MPBA.
Separations 10 00004 g004
Figure 5. Thermogravimetric analysis curves of phenylboronic-acid-modified magnetic nanoparticles.
Figure 5. Thermogravimetric analysis curves of phenylboronic-acid-modified magnetic nanoparticles.
Separations 10 00004 g005
Figure 6. Magnetization curves of phenylboronic-acid-modified magnetic nanoparticles at room temperature.
Figure 6. Magnetization curves of phenylboronic-acid-modified magnetic nanoparticles at room temperature.
Separations 10 00004 g006
Figure 7. Adsorption isothermal curves of Fe3O4@MPTES@VPBA.
Figure 7. Adsorption isothermal curves of Fe3O4@MPTES@VPBA.
Separations 10 00004 g007
Figure 8. Scatchard fitting curve of Fe3O4@MPTES@VPBA (a,c) and Fe3O4@MPTES (b,d) for the adsorption of adenosine (a,b) and o-dihydroxybenzene (c,d).
Figure 8. Scatchard fitting curve of Fe3O4@MPTES@VPBA (a,c) and Fe3O4@MPTES (b,d) for the adsorption of adenosine (a,b) and o-dihydroxybenzene (c,d).
Separations 10 00004 g008
Figure 9. Kinetics of adsorption curves of Fe3O4@MPTES@VPBA.
Figure 9. Kinetics of adsorption curves of Fe3O4@MPTES@VPBA.
Separations 10 00004 g009
Figure 10. HPLC chromatograms of the mixed solution of adenosine and deoxyadenosine (a) and the chromatograms of the eluent for Fe3O4@MPS@MPBA (b), Fe3O4@MPTES@AAPBA (c) and Fe3O4@MPTES@VPBA (d) after being adsorbed in the mixture solution. Peak area ratio of adenosine to deoxyadenosine: 1:0.26 (b), 1:0.29 (c), 1:0.27 (d).
Figure 10. HPLC chromatograms of the mixed solution of adenosine and deoxyadenosine (a) and the chromatograms of the eluent for Fe3O4@MPS@MPBA (b), Fe3O4@MPTES@AAPBA (c) and Fe3O4@MPTES@VPBA (d) after being adsorbed in the mixture solution. Peak area ratio of adenosine to deoxyadenosine: 1:0.26 (b), 1:0.29 (c), 1:0.27 (d).
Separations 10 00004 g010
Figure 11. Chromatograms of mixed solution of o-dihydroxybenzene, m-dihydroxybenzene and p-dihydroxybenzene (a) and the chromatograms of the eluent for Fe3O4@MPS@MPBA (b), Fe3O4@MPTES@AAPBA (c) and Fe3O4@MPTES@VPBA (d) after being adsorbed in the mixture solution. Peak area ratio of o-dihydroxybenzene to m- and p-: 1:0.12: 0.06 (b), 1:0.26: 0.17 (c), 1:0.21:0.14 (d).
Figure 11. Chromatograms of mixed solution of o-dihydroxybenzene, m-dihydroxybenzene and p-dihydroxybenzene (a) and the chromatograms of the eluent for Fe3O4@MPS@MPBA (b), Fe3O4@MPTES@AAPBA (c) and Fe3O4@MPTES@VPBA (d) after being adsorbed in the mixture solution. Peak area ratio of o-dihydroxybenzene to m- and p-: 1:0.12: 0.06 (b), 1:0.26: 0.17 (c), 1:0.21:0.14 (d).
Separations 10 00004 g011
Table 1. KD of different Fe3O4@PBA for the adsorption of adenosine and o-dihydroxybenzene.
Table 1. KD of different Fe3O4@PBA for the adsorption of adenosine and o-dihydroxybenzene.
AdsorbateAdsorbentKD (mg/mL)
High-Affinity Binding SitesLow-Affinity Binding Sites
AdenosineFe3O4@MPTES@VPBA0.131.29
Fe3O4@MPTES@ AAPBA0.051.95
Fe3O4@MPTES3.68
Fe3O4@MPS@MPBA0.051.33
Fe3O4@MPS3.45
o-DihydroxybenzeneFe3O4@MPTES@VPBA0.412.50
Fe3O4@MPTES@ AAPBA0.332.64
Fe3O4@MPTES5.15
Fe3O4MPS@MPBA0.132.73
Fe3O4@MPS4.17
Table 2. Absorption capacity (Qe) and selectivity factor (α) of different adsorbates by Fe3O4@PBA nanoparticles.
Table 2. Absorption capacity (Qe) and selectivity factor (α) of different adsorbates by Fe3O4@PBA nanoparticles.
AdsorbateFe3O4@MPTES@VPBAFe3O4@MPTES@AAPBAFe3O4@MPS@MPBA
Qe (mg/g)αQe (mg/g)αQe (mg/g)α
Adenosine91.9-92.7-94.6-
Deoxyadenosine18.54.9719.14.8517.35.47
o-Dihydroxybenzene80.6-79.3-83.1
m-Dihydroxybenzene34.62.3336.42.1828.52.92
p-Dihydroxy-benzene31.12.5932.62.4326.23.17
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, H.; Zhang, J.; Duan, A.; Wang, B.; Xie, S.; Yuan, L. Facile Preparation of Phenyboronic-Acid-Functionalized Fe3O4 Magnetic Nanoparticles for the Selective Adsorption of Ortho-Dihydroxy-Containing Compounds. Separations 2023, 10, 4. https://doi.org/10.3390/separations10010004

AMA Style

Zhou H, Zhang J, Duan A, Wang B, Xie S, Yuan L. Facile Preparation of Phenyboronic-Acid-Functionalized Fe3O4 Magnetic Nanoparticles for the Selective Adsorption of Ortho-Dihydroxy-Containing Compounds. Separations. 2023; 10(1):4. https://doi.org/10.3390/separations10010004

Chicago/Turabian Style

Zhou, Hongmei, Junhui Zhang, Aihong Duan, Bangjin Wang, Shengming Xie, and Liming Yuan. 2023. "Facile Preparation of Phenyboronic-Acid-Functionalized Fe3O4 Magnetic Nanoparticles for the Selective Adsorption of Ortho-Dihydroxy-Containing Compounds" Separations 10, no. 1: 4. https://doi.org/10.3390/separations10010004

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop