Next Article in Journal
Influence of Graphene Oxide Concentration and Ultrasonication Energy on Fracture Behavior of Nano-Reinforced Cement Pastes
Previous Article in Journal
Twin Peaks: Interrogating Otolith Pairs to See Whether They Keep Their Stories Straight
Previous Article in Special Issue
Synthesis, Characterisation, and Applications of TiO and Other Black Titania Nanostructures Species (Review)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of Iron–Sillenite for Application as an XRD/MRI Dual-Contrast Agent

1
Institute of Chemistry, Vilnius University, Naugarduko St. 24, 03225 Vilnius, Lithuania
2
Graduate Institute of Nanomedicine and Medical Engineering, Taipei Medical University, No. 301, Yuantong Road, Zhonghe District, New Taipei City 235603, Taiwan
3
Graduate Institute of Biomedical Materials and Tissue Engineering, College of Biomedical Engineering, Taipei Medical University, Taipei 11052, Taiwan
4
Division of Gastroenterology and Hepatology, Department of Internal Medicine, Shuang Ho Hospital, Taipei Medical University, Zhongzheng Road, Zhonghe, Taipei 23561, Taiwan
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(8), 706; https://doi.org/10.3390/cryst14080706
Submission received: 22 June 2024 / Revised: 23 July 2024 / Accepted: 1 August 2024 / Published: 5 August 2024

Abstract

:
In the present work, iron–sillenite (Bi25FeO40) was synthesized using a simple solid-state reaction method and characterized. The effects of the synthesis conditions on the phase purity of Bi2O3/Fe3O4, morphological features, and possible application as an XRD/MRI dual-contrast agent were investigated. For the synthesis, the stoichiometric amounts of Bi2O3 and Fe3O4 were mixed and subsequently milled in a planetary ball mill for 10 min with a speed of 300 rpm. The milled mixture was calcined at various temperatures (550 ° C, 700 ° C, 750 ° C, 800 ° C, and 850 ° C) for 1 h in air at a heating rate of 5 °C/min. For phase identification, powder X-ray diffraction (XRD) analysis was performed and infrared (FTIR) spectra were recorded. The surface morphology of synthesized samples was studied by field-emission scanning electron microscopy (FE-SEM). For the radiopacity measurements, iron–sillenite specimens were synthesized at different temperatures and mixed with different amounts of BaSO4 and Laponite solution. It was demonstrated that iron–sillenite Bi25FeO40 possessed sufficient radiopacity and could be a potential candidate to meet the requirements of its application as an XRD/MRI dual-contrast agent.

1. Introduction

The development of biomedical imaging techniques like X-ray, computed tomography (CT), magnetic resonance imaging (MRI), positron emission tomography (PET), digital mammography (DM), optical imaging (OI), and ultrasound (US) is of great importance to obtain comprehensive information about tissues and organs [1]. Currently, various multimodal-contrast-based agents imaging modalities, such as X-ray/US [2], CT/MRI [3], OI/MRI [4], PET/CT [5], and PET/MRI [6], are successfully applied to obtain high sensitivity and anatomic resolution in clinical diseases diagnosis.
Magnetic resonance imaging (MRI) provides detailed images of brain, spine, joint, and soft-tissue examinations [7]. Iron oxide nanoparticles like Fe2O3, with their inherent magnetism, biocompatibility, and flexibility of engineering, are ideal candidates for MRI and multimodal imaging [8]. Radiopacity is essential for tracking medical devices in X-ray images during procedures. Bismuth oxide (Bi2O3) is widely used for dental filling materials due to its acceptable radiopacity to distinguish it from surrounding anatomic structures [9]. A modified Bi2O3/ZrO2 radiopacifier was proposed to improve its biocompatibility [10].
Solid-state synthesis has been conventionally used for inorganic compounds, but there is also a growing interest in solvent-free organic synthesis to design new materials [11]. Solid-state reaction syntheses offer several advantages related to their adaptability with multi-elemental phases from a mixture of solid starting materials in direct reactions at high temperatures [12].
Sillenite-type members of the bismuth ferrite family have demonstrated outstanding potential as novel photocatalysts in environmental remediation [13]. It was shown recently that iron–sillenite (Bi25FeO40) can be successfully used as a supercapacitor candidate and for dye-sensitized solar cells [14]. The structural, optical, dielectric, catalytic, and magnetic properties of Bi25FeO40 can be controlled by changing particle size, by doping with other metals, or by developing composites [15,16]. For example, with increasing Cr concentration in Cr-doped Bi25FeO40 (Bi25Fe1-xCrxO40, x = 0–0.5), the shape of particles has changed from microcubes (x = 0) to microspheres (x = 0.3) owing to the lattice expansion [17]. The Sb-doped Bi25FeO40 showed significant electrocatalytic activity [18]. The composite with SiO2 (Bi25FeO40/SiO2) had quite different microstructural, optical, and magnetic properties and an improved specific absorption rate of the nanoparticles [19].
Firstly, iron–sillenite was isolated as the impurity phase during the preparation of perovskite BiFeO3 (BFO) or its derivatives [20]. For example, during the synthesis of Bi1-xLaxFeO3 by the co-precipitation method, single-phase BFO phase formed when x = 0.3. With an increasing amount of lanthanum, the impurity Bi25FeO40 phase was detected in the synthesis products [21]. When the sol–gel–combustion synthesis method was applied for preparation of BFO, the pure phase was obtained only at a low heating rate and annealing temperatures between 500 °C and 600 °C. It was determined that, above 600 °C, the BiFeO3 gradually decomposed to Bi25FeO40 and Bi2Fe4O9 [22]. Goldman et al. [23] also observed that sillenite formed as intermediate product during synthesis of BFO using a hydrothermal approach. Similar results have been published by Sansom et al. [24] and Yang et al. [25]. Later, the monophasic Bi25FeO40 was synthesized using hydrothermal [26,27,28], combustion [26], molten salts [29], and mechanical [30] synthesis methods. The final products obtained by different synthesis approaches showed different surface morphology and physical properties.
Different composites with Bi25FeO40 have been also synthesized and investigated. The Bi25FeO40–graphene composite photocatalyst exhibited higher catalytic activity, was superparamagnetic, and can be readily recovered in an external magnetic field [31]. The Bi/Bi25FeO40-C even exhibited higher catalytic efficiency [32]. It was demonstrated that the sillenite–graphene oxide nanocomposite is promising for improving the magnetic and optical properties for potential technological applications [33,34]. Silver phosphate/sillenite bismuth ferrite/graphene oxide (Ag3PO4/Bi25FeO40/GO) nanocomposite has been successfully fabricated as well [35]. In the work by [36], a Bi25FeO40-Fe3O4-Fe2O3 composite was prepared directly through the solid-state reaction process. The results of the investigation indicated that these composite samples have photocatalytic properties which can be easily recycled by magnetic separation.
The heterojunction nanostructures of bismuth iron oxides showed enhanced photocatalytic and other properties. A Bi25FeO40/Bi2Fe4O9 photocatalyst has been synthesized and evaluated as a visible-light responsive catalyst for the degradation of Rhodamine B [37]. A heterojunction-type charge transfer mechanism interpreting the enhanced photocatalytic activities was proposed and discussed in this study. The strategies of BiFeO3/Bi25FeO40 heterojunction construction, temperature, and morphology controlling and Fe3+ doping allowed precise regulation of the band gap structure of bismuth ferrite [38,39]. These findings promoted the application of BiFeO3 in photocatalytic and other redox reactions.
Magnetic resonance imaging (MRI) has become one of the most prominent and widely adapted diagnostic imaging methods in the realm of clinical practice and biomedical research because of its superior spatial resolution and 3D tomographic images with anatomical details. To date, only a few nanosystems have been investigated as T1 MRI-CT dual-contrast agents. The summarized results suggest that D-glucuronic acid-coated Gd(IO3)3 center dot [40], low-magnetization magnetite nanocubes [41], Eu-doped iron oxide nanoparticles [42], functionalized Fe3O4 composites [43], or (Fe3O4/γ-Fe2O3) nanoparticles coated by gadolinium–diethylenetriaminepentaacetic acid [44] could be potential T1 MRI-CT dual-contrast reagents. However, these dual-contrast agents have a relatively low sensitivity compared with other imaging methods. Moreover, most of them suffer from low relaxation and contrast efficiency, which hampers their application in clinical diagnosis.
In this study, we aim to fill this gap. Iron–sillenite (Bi25FeO40) was synthesized using a simple solid-state reaction method and characterized. The effects of the synthesis conditions on the phase purity of Bi2O3/Fe3O4, morphological features, and possible application as an XRD/MRI dual-contrast agent were investigated.

2. Materials and Methods

2.1. Synthesis

Iron–sillenite (Bi25FeO40) was synthesized using a simple solid-state reaction method. Stoichiometric amounts of Bi2O3 (Alfa Aesar, Haverhill MA, USA, 99.9%) and Fe3O4 (Sigma-Aldrich, St. Louis, MO, USA, 97.0%) were mixed and subsequently milled in a planetary ball mill for 10 min with a speed of 300 rpm. Ball milling was performed using a bench-top planetary ball mill (Retsch PM100, Haan, Germany). The conditions involved loading a 10 g sample into a 50 mL alumina grinding jar and adding 10 alumina grinding balls (each with a diameter of 10 mm). The milled mixture was calcined at various temperatures (550 °C, 700 °C, 750 °C, 800 °C, and 850 °C) for 1 h in air at a heating rate of 5 °C/min. Finally, the synthesized products were ground in an agate mortar. A simplified scheme of the synthesis route of iron–sillenite is presented in Figure 1.

2.2. Characterization

For phase identification at ambient temperature, the XRD data were collected at 20–70° 2 θ range (step width of 0.2°, scan speed 3.33°/min) using Ni-filtered Cu Kα1 (λ = 1.54184 Å) radiation on a Bruker D2 PHASER diffractometer. Infrared (FTIR) spectra were recorded in the range of 4000−400 cm−1 employing a Bruker ALPHA ATR spectrometer. In order to study the morphology of the samples, a field-emission scanning electron microscope (FE-SEM) Hitachi SU-70 (Tokyo, Japan) was used. The radiopacity of samples was measured using a dental X-ray system (VX-65, Shanghai, China) and X-ray images recorded by an occlusal radiographic imaging plate (Kodak CR. Los Angeles, CA, USA). The mean grayscale values of each step of the aluminum step wedge (from 2 to 16 mm in 2 mm increments) and the specimens were measured and analyzed using imaging processing software ImageJ 1.39f (Wayne Rasband). For the radiopacity measurements, iron–sillenite samples were synthesized at different temperatures and mixed with different amounts of BaSO4 (Alfa Aesar, Haverhill MA, USA, 99.0%) and Laponite solution (Sigma-Aldrich, St. Louis, MO, USA, LiMgNaO6Si2, 1% solution).

3. Results and Discussion

The XRD patterns of Bi25FeO40 synthesis products obtained at 700 °C, 750 °C, and 800 °C are presented in Figure 2.
As can be seen from the XRD diffraction patterns of synthesis products obtained at 750 °C, all the peaks match very well with the standard XRD data of Bi25FeO40 (PDF #96-900-5813). Also, iron–sillenite was the main crystalline phase in the products annealed at slightly lower or higher temperatures. However, unreacted Bi2O3 was determined in the XRD pattern of the sample obtained at 700 °C, while perovskite BiFeO3 and some unidentified phase formed at 800 °C. Evidently, the solid-state reaction performed at 550 °C was not complete and the reaction mixture annealed at 850 °C was already melted with the formation of multiphasic product (see Figure S1). Thus, the apparently optimal temperature for solid-state reaction synthesis of monophasic Bi25FeO40 is 750 °C.
FTIR spectra of the products obtained at 700 °C, 750 °C, and 800 °C are presented in Figure 3. The FTIR range of 1250–400 cm−1 was chosen as representative since the main bands attributed to Bi25FeO40 synthesis products can be observed in this region. As can be seen, only stretching modes of metal–oxygen (M–O) vibrations can be observed at approximately 820 cm−1, and in the range of 625–430 cm−1 in the FTIR spectra of all three samples. These absorption bands correspond to the vibration of the Fe–O, Bi–O, or Bi–O–Fe bonds in the crystalline lattice of Bi25FeO40 [45,46,47,48]. Such results are in a good agreement with the XRD data.
The morphology of the Bi25FeO40 synthesis products obtained at 700 °C, 750 °C, and 800 °C was almost identical. However, the samples sintered at 550 °C and 850 °C showed quite different morphological features. The SEM micrographs of the representative samples are shown in Figure 4.
As can be seen, the particle shape varied from plate-like, spherical to multishaped and flower-like structures by changing the synthesis temperature from 550 °C to 850 °C. The samples fabricated at 700–800 °C were composed of spherical shapes and grew into each other’s particles at 1–2 µm in size. It is worth noting that the narrow particle size distribution was achieved even though the synthesis was performed by the solid-state reaction method. The results of EDX analysis confirmed the molar ratio of Bi and Fe in synthesized Bi25FeO40. The EDX spectrum and color mapping, along with the SEM image of the representative sample, is provided in Figure S2.
Radiopacity is a key factor to verify the efficiency of a radiocontrast agent in imaging [49]. The synthesized samples were also characterized by the radiopacity measurements. For the radiopacity measurements, the Bi25FeO40 synthesized at 750 °C was mixed with different amount of BaSO4 and Laponite solution. Initially, the control reference samples were prepared from BaSO4 and Laponite using different ratios of constituents (see Table 1). The stability of the reference samples was checked visually. It is evident from digital photos presented in Figure 5 that the color of different mixtures of BaSO4 and Laponite were stable for 12 h. The aluminum step wedge (99.5% Al) was used as an internal standard for measuring the equivalent radiopacity of different materials [50,51]. The results of the radiopacity of reference samples with composition presented in Table 1 are shown in Table 2. The grayscale value corresponds to the attenuation of the material. The measured grayscale value for each reference composition and aluminum corresponds to the amount of attenuation. The regression parameter R2 for the investigated systems varied in the range of 0.9948–0.9977.
As can be seen from Table 2, the radiopacity values for the system A2, B2, and C2 were higher in comparison with A1, B1, and C1. Therefore, the BaSO4 and Laponite samples from series A2, B2, and C2 were selected for the investigation of the radiopacity of the iron–sillenite sample synthesized at 750 °C using a solid-state reaction method. These samples are marked A3, B3, and C3, respectively. Again, the samples with iron–sillenite showed excellent stability over time (see Figure 6).
Visual examination of the radiographic images (Figure 7) revealed that all three A3, B3, and C3 specimens were homogeneous [51,52]. Evidently, the samples are free of radiolucent and radiopaque inclusions.
To quantify the radiopacity, an Al wedge was again placed beside the iron–sillenite material during X-ray image acquisition and the grayscale values of the material of interest along with the step wedge were digitally analyzed. The radiopacity of the specimen was then referenced to the thickness of aluminum and expressed as the equivalent aluminum thickness (mm Al) [53]. The calibration curves were plotted using best-fit logarithmic regression analysis for the selected data. The equivalent in thickness of aluminum for each material was calculated from the calibration curves. The average magnitude of the mean regression residuals was 0.040 mm of aluminum and the maximum regression residual was 0.125 mm of aluminum The residuals were random with respect to radiopacity, indicating that no major non-linearity was present. The results’ radiopacity of iron–sillenite Bi25FeO40 specimens obtained are presented in Figure 8 and Table 3.
In conclusion, the radiopacities of the tested B3 and C3 materials were greater than 0.5 mm and greater than the same thickness of aluminum. Therefore, the solid-state reaction-derived iron–sillenite Bi25FeO40 possessed sufficient radiopacity and could be a potential candidate to meet the requirements of its application as an XRD/MRI dual-contrast agent.

4. Conclusions

In this study, iron–sillenite (Bi25FeO40) was synthesized using a simple solid-state reaction method and characterized. The effects of the synthesis conditions on the phase purity of Bi2O3/Fe3O4, morphological features, and possible application as an XRD/MRI dual-contrast agent were investigated. To obtain the iron–sillenite phase, the mixture of starting reagents was calcined at various temperatures (550 °C, 700 °C, 800 °C, and 850 °C). The powder XRD analysis data showed that the synthesis product obtained at 750 °C was monophasic Bi25FeO40. Also, iron–sillenite was the main crystalline phase in the products annealed at slightly lower or higher temperatures (700 °C and 800 °C, respectively). However, unreacted Bi2O3 was determined in the XRD pattern of the sample obtained at 700 °C, while perovskite BiFeO3 and some unidentified phase formed at 800 °C. Evidently, the solid-state reaction performed at 550 °C was not complete and the reaction mixture annealed at 850 °C was already melted with the formation of multiphasic product. FTIR spectroscopy results were in a good agreement with the XRD data. The samples fabricated at 700–800 °C were composed of spherical shapes and grew into each other’s particles at 1–2 µm in size. The synthesized samples were also characterized by the radiopacity measurements. For the radiopacity measurements, the Bi25FeO40 synthesized at 750 °C was mixed with different amount of BaSO4 and Laponite solution. Visual examination of the radiographic images revealed that the specimens were homogeneous. The radiopacities of the best compositions were greater than 0.5 mm and greater than the same thickness aluminum. Therefore, the solid-state reaction-derived iron–sillenite Bi25FeO40 possessed sufficient radiopacity and could be a potential candidate to meet the requirements of its application as an XRD/MRI dual-contrast agent.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst14080706/s1, Figure S1: XRD patterns of Bi25FeO40 synthesis products obtained at 550 and 850 °C along with standard XRD pattern of starting materials; Figure S2: The EDX spectrum (top) and colour mapping along with SEM images (bottom) of Bi25FeO40 synthesis products obtained at 700 °C.

Author Contributions

Conceptualization, J.-C.Y.; methodology, A.K., L.-Y.C. and A.Z., investigation, Y.-T.W., P.-W.L., D.V. and I.G., resources, J.-C.Y. and A.K., data curation, L.-Y.C., Y.-T.W., D.V., I.G. and A.Z., writing—original draft preparation, A.K. and J.-C.Y., writing—review and editing, A.K. and J.-C.Y., supervision, A.K. and J.-C.Y.; funding acquisition, A.K. and J.-C.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research received funding by NSTC112-2221-E-038-013 from Taipei Medical University and Vilnius University.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hussain, S.; Mubeen, I.; Ullah, N.; Shahab, S.; Shah, U.D.; Khan, B.A.; Zahoor, M.; Ullah, R.; Khan, F.A.; Sultan, M.A. Modern Diagnostic Imaging Technique Applications and Risk Factors in the Medical Field: A Review. Biomed. Res. Int. 2022, 2022, 5164970. [Google Scholar] [CrossRef] [PubMed]
  2. Aimacana, C.M.C.; Perez, D.A.Q.; Pinto, S.R.; Debut, A.; Attia, M.F.; Santos-Oliveira, R.; Whitehead, D.C.; Terencio, T.; Alexis, F.; Dahoumane, S.A. Polytetrafluoroethylene-like Nanoparticles as a Promising Contrast Agent for Dual Modal Ultrasound and X-ray Bioimaging. ACS Biomater. Sci. Eng. 2021, 7, 1181–1191. [Google Scholar] [CrossRef] [PubMed]
  3. Ventura, M.; Sun, Y.; Rusu, V.; Laverman, P.; Borm, P.; Heerschap, A.; Oosterwijk, E.; Boerman, O.C.; Jansen, J.A.; Walboomers, X.F. Dual contrast agent for computed tomography and magnetic resonance hard tissue imaging. Tiss. Eng. C Methods 2013, 19, 405–416. [Google Scholar] [CrossRef] [PubMed]
  4. Zhu, T.; Ma, X.; Chen, R.; Ge, Z.; Xu, J.; Shen, X.; Jia, L.; Zhou, T.; Luo, Y.; Ma, T. Using fluorescently-labeled magnetic nanocomposites as a dual contrast agent for optical and magnetic resonance imaging. Biomater. Sci. 2017, 5, 1090–1100. [Google Scholar] [CrossRef] [PubMed]
  5. Farwell, M.D.; Pryma, D.A.; Mankoff, D.A. PET/CT imaging in cancer: Current applications and future directions. Cancer 2014, 120, 3433–3445. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, Y.; García-Gabilondo, M.; Grayston, A.; Feiner, I.V.J.; Anton-Sales, I.; Loiola, R.A.; Llop, J.; Ramos-Cabrer, P.; Barba, I.; Garcia-Dorado, D.; et al. PLGA protein nanocarriers with tailor-made fluorescence/MRI/PET imaging modalities. Nanoscale 2020, 12, 4988–5002. [Google Scholar] [CrossRef] [PubMed]
  7. Westbrook, C.; Talbot, J. MRI in Practice; Wiley: New York, NY, USA, 2018; p. 416. [Google Scholar]
  8. El-Hammadi, M.M.; Arias, J.L. Iron oxide-based multifunctional nanoparticulate systems for biomedical applications: A patent review (2008–present). Expert Opin. Ther. Pat. 2015, 25, 691–709. [Google Scholar] [CrossRef] [PubMed]
  9. Hsieh, S.-C.; Teng, N.-C.; Lin, C.-K.; Lee, P.-Y.; Ji, D.-Y.; Chen, C.-C.; Ke, E.-S.; Lee, S.-Y.; Yang, J.-C. A Novel Accelerator for Improving the Handling Properties of Dental Filling Materials. J. Endod. 2009, 35, 1292–1295. [Google Scholar] [CrossRef] [PubMed]
  10. Chen, C.-C.; Hsieh, S.-C.; Teng, N.-C.; Kao, C.-K.; Lee, S.-Y.; Lin, C.-K.; Yang, J.-C. Radiopacity and cytotoxicity of Portland cement containing zirconia doped bismuth oxide radiopacifiers. J. Endod. 2014, 40, 251–254. [Google Scholar] [CrossRef] [PubMed]
  11. Smart, L.E.; Moore, E.A. Solid State Chemistry: An Introduction, 4th ed.; Taylor & Francis: Boca Raton, FL, USA, 2012. [Google Scholar]
  12. Kumar, A.; Dutta, S.; Kim, S.; Kwon, T.; Patil, S.S.; Kumari, N.; Jeevanandham, S.; Lee, I.S. Solid-State Reaction Synthesis of Nanoscale Materials: Strategies and Applications. Chem. Rev. 2022, 122, 12748–12863. [Google Scholar] [CrossRef] [PubMed]
  13. Sharmin, F.; Basith, M.A. Simple Low Temperature Technique to Synthesize Sillenite Bismuth Ferrite with Promising Photocatalytic Performance. ACS Omega 2022, 7, 34901–34911. [Google Scholar] [CrossRef] [PubMed]
  14. Kumar, A.M.; Ragavendran, V.; Mayandi, J.; Ramachandran, K.; Jayakumar, K. Influence of PVP on Bi25FeO40 microcubes for Supercapacitors and Dye-Sensitized Solar Cells applications. J. Mater. Sci. Mater. Electr. 2022, 33, 9512–9524. [Google Scholar] [CrossRef]
  15. Jebari, H.; Tahiri, N.; Boujnah, M.; El Bounagui, O.; Boudad, L.; Taibi, M.; Ez-Zahraouy, H. Structural, optical, dielectric, and magnetic properties of iron-sillenite Bi25FeO40. Appl. Phys. 2022, 128, 842. [Google Scholar] [CrossRef]
  16. Nayak, A.K.; Gopalakrishnan, T. Phase- and Crystal Structure-Controlled Synthesis of Bi2O3, Fe2O3, and BiFeO3 Nanomaterials for Energy Storage Devices. ACS Appl. Nanomater. 2022, 5, 14663–14676. [Google Scholar] [CrossRef]
  17. Xiong, Z.W.; Cao, L.H. Tailoring morphology, enhancing magnetization and photocatalytic activity via Cr doping in Bi25FeO40. J. Alloys Compd. 2019, 773, 828–837. [Google Scholar] [CrossRef]
  18. Zhang, Y.; Cao, S.H.; Liang, C.; Shen, J.M.; Chen, Y.Q.; Feng, Y.C.; Chen, H.; Liu, R.; Jiang, F. Electrocatalytic performance of Sb-modified Bi25FeO40 for nitrogen fixation. J. Colloid Interf. Sci. 2021, 593, 335–344. [Google Scholar] [CrossRef] [PubMed]
  19. Juwita, E.; Sulistiani, F.A.; Darmawan, M.Y.; Istiqomah, N.I.; Suharyadi, E. Microstructural, optical, and magnetic properties and specific absorption rate of bismuth ferrite/SiO2 nanoparticles. Mater. Res. Express 2022, 9, 076101. [Google Scholar] [CrossRef]
  20. Wu, L.; Dong, C.H.; Chen, H.; Yao, J.L.; Jiang, C.J.; Xue, D.S. Hydrothermal Synthesis and Magnetic Properties of Bismuth Ferrites Nanocrystals with Various Morphology. J. Am. Ceram. Soc. 2012, 95, 3922–3927. [Google Scholar] [CrossRef]
  21. Yotburut, B.; Yamwong, T.; Thongbai, P.; Maensiri, S. Synthesis and characterization of coprecipitation-prepared La-doped BiFeO3 nanopowders and their bulk dielectric properties. Jpn. J. Appl. Phys. 2014, 53, 06JG13. [Google Scholar] [CrossRef]
  22. Koeferstein, R. Synthesis, phase evolution and properties of phase-pure nanocrystalline BiFeO3 prepared by a starch-based combustion method. J. Alloys Compd. 2014, 590, 324–330. [Google Scholar] [CrossRef]
  23. Goldman, A.R.; Fredricks, J.L.; Estroff, L.A. Exploring reaction pathways in the hydrothermal growth of phase-pure bismuth ferrites. J. Cryst. Growth 2017, 468, 104–109. [Google Scholar] [CrossRef]
  24. Sansom, G.; Rattanakam, R.; Jettanasen, J. Effects of Scaling Up on the Phase Evolution of Microcrystalline Bismuth Ferrite during Hydrothermal Process. E-J. Surf. Sci. Nanotechnol. 2022, 20, 85–89. [Google Scholar] [CrossRef]
  25. Yang, X.; Xu, G.; Ren, Z.H.; Weng, W.J.; Du, P.Y.; Shen, G.; Han, G.R. Effect of PVA Adding Amount on Phase-Controlled Synthesis and Morphology Evolution of the Bismuth Ferrite by Assisted Hydrothermal Reaction Route. Rare Met. Mater. Eng. 2012, 41, 247–249. [Google Scholar]
  26. Koeferstein, R.; Buttlar, T.; Ebbinghaus, S.G. Investigations on Bi25FeO40 powders synthesized by hydrothermal and combustion-like processes. J. Solid State Chem. 2014, 217, 50–56. [Google Scholar] [CrossRef]
  27. Ji, W.D.; Li, M.M.; Zhang, G.; Wang, P. Controlled synthesis of Bi25FeO40 with different morphologies: Growth mechanism and enhanced photo-Fenton catalytic properties. Dalton Trans. 2017, 46, 10586–10593. [Google Scholar] [CrossRef] [PubMed]
  28. Kumar, A.M.; Ragavendran, V.; Mayandi, J.; Ramachandran, K.; Jayakumar, K. Phase dependent electrochemical characteristics of bismuth ferrite: A bifunctional electrocatalyst for Supercapacitors and Dye-Sensitized Solar Cells. Colloids Surf. A Physicochem. Eng. Asp. 2023, 656, 130529. [Google Scholar] [CrossRef]
  29. Ren, L.; Lu, S.Y.; Fang, J.Z.; Wu, Y.; Chen, D.Z.; Huang, L.Y.; Chen, Y.F.; Cheng, C.; Liang, Y.; Fang, Z.Q. Enhanced degradation of organic pollutants using Bi25FeO40 microcrystals as an efficient reusable heterogeneous photo-Fenton like catalyst. Catal. Today 2017, 281, 656–661. [Google Scholar] [CrossRef]
  30. Zou, W.J.; Dong, J.T.; Ji, M.X.; Wang, B.; Li, Y.J.; Yin, S.; Li, H.M.; Xia, J.X. Synthesis of Bi25FeO40 Nanoparticles with Oxygen Vacancies via Ball Milling for Fenton Oxidation of Tetracycline Hydrochloride and Reduction of Cr(VI). ACS Appl. Nano Mater. 2023, 6, 4309–4318. [Google Scholar] [CrossRef]
  31. Sun, A.W.; Chen, H.; Song, C.Y.; Jiang, F.; Wang, X.; Fu, Y.S. Magnetic Bi25FeO40-graphene catalyst and its high visible-light photocatalytic performance. RSC Adv. 2013, 3, 4332–4340. [Google Scholar] [CrossRef]
  32. Li, F.H.; Zhou, J.K.; Gao, C.J.; Qiu, H.X.; Gong, Y.L.; Gao, J.H.; Liu, Y.; Gao, J.P. A green method to prepare magnetically recyclable Bi/Bi25FeO40-C nanocomposites for photocatalytic hydrogen generation. Appl. Surf. Sci. 2020, 521, 146342. [Google Scholar] [CrossRef]
  33. Jalil, M.A.; Chowdhury, S.S.; Sakib, M.A.; Yousuf, S.M.E.H.; Ashik, E.K.; Firoz, S.H.; Basith, M.A. Temperature-dependent phase transition and comparative investigation on enhanced magnetic and optical properties between sillenite and perovskite bismuth ferrite-rGO nanocomposites. J. Appl. Phys. 2017, 122, 084902. [Google Scholar] [CrossRef]
  34. Basith, M.A.; Ahsan, R.; Zarin, I.; Jalil, M.A. Enhanced photocatalytic dye degradation and hydrogen production ability of Bi25FeO40-rGO nanocomposite and mechanism insight. Sci. Rep. 2018, 8, 11090. [Google Scholar] [CrossRef] [PubMed]
  35. Huang, Y.; Zhang, X.Y.; Zhu, G.X.; Gao, Y.; Cheng, Q.F.; Cheng, X.W. Synthesis of silver phosphate/sillenite bismuth ferrite/graphene oxide nanocomposite and its enhanced visible light photocatalytic mechanism. Separ. Purif. Technol. 2019, 215, 490–499. [Google Scholar] [CrossRef]
  36. de Gois, M.M.; Araujo, W.P.; da Silva, R.B.; da Luz, G.E.; Soares, J.M. Bi25FeO40-Fe3O4-Fe2O3 composites: Synthesis, structural characterization, magnetic and UV-visible photocatalytic properties. J. Alloys Compd. 2019, 785, 598–602. [Google Scholar] [CrossRef]
  37. Wang, G.M.; Cheng, D.; He, T.C.; Hu, Y.Y.; Deng, Q.R.; Mao, Y.W.; Wang, S.G. Enhanced visible-light responsive photocatalytic activity of Bi25FeO40/Bi2Fe4O9 composites and mechanism investigation. J. Mater. Sci. Mater. Electron. 2019, 30, 10923–10933. [Google Scholar] [CrossRef]
  38. Wang, Y.F.; Xu, C.X.; Yan, L.; Li, J. Synthesis of BiFeO3/Bi25FeO40 heterojunction structure and precise adjustment of forbidden band width. Mater. Chem. Phys. 2023, 305, 127935. [Google Scholar] [CrossRef]
  39. Xu, C.X.; Wang, Y.F.; Wang, Q.; Li, J.; Yan, L. Phase transformation and heterojunction nanostructures of bismuth iron oxide. J. Mater. Sci. Mater. Electron. 2023, 34, 2236. [Google Scholar] [CrossRef]
  40. Lee, E.J.; Heo, W.C.; Park, J.W.; Chang, Y.; Bae, J.-E.; Chae, K.S.; Kim, T.J.; Park, J.A.; Lee, G.H. D-Glucuronic Acid Coated Gd(IO3)3·2H2O Nanomaterial as a Potential T1 MRI-CT Dual Contrast Agent. Eur. J. Inorg. Chem. 2013, 16, 2858–2866. [Google Scholar] [CrossRef]
  41. Sharma, V.K.; Alipour, A.; Soran-Erdem, Z.; Aykut, Z.G.; Demir, H.V. Highly monodisperse low-magnetization magnetite nanocubes as simultaneous T1-T2 MRI contrast agents. Nanoscale 2015, 7, 10519–10526. [Google Scholar] [CrossRef] [PubMed]
  42. Park, J.C.; Lee, G.T.; Kim, H.-K.; Sung, B.; Lee, Y.; Kim, M.; Chang, Y.; Seo, J.H. Surface Design of Eu-Doped Iron Oxide Nanoparticles for Tuning the Magnetic Relaxivity. ACS Appl. Mater. Interf. 2018, 10, 25080–25089. [Google Scholar] [CrossRef] [PubMed]
  43. Illert, P.; Waengler, B.; Waengler, C.; Zoellner, F.; Uhrig, T.; Litau, S.; Pretze, M.; Roeder, T. Functionalizable composite nanoparticles as a dual magnetic resonance imaging/computed tomography contrast agent for medical imaging. J. Appl. Polym. Sci. 2019, 136, 47571. [Google Scholar] [CrossRef]
  44. Eguia-Eguia, S.I.; Gildo-Ortiz, L.; Perez-Gonzalez, M.; Tomas, S.A.; Arenas-Alatorre, J.A.; Santoyo-Salazar, J. Magnetic domains orientation in (Fe3O4/γ-Fe2O3) nanoparticles coated by Gadolinium-diethylenetriaminepentaacetic acid (Gd3+-DTPA). Nano Express 2021, 2, 020019. [Google Scholar] [CrossRef]
  45. Kun Yang, K.; Peng, H.; Wen, Y.; Li, N. Re-examination of characteristic FTIR spectrum of secondary layer in bilayer oleic acid-coated Fe3O4 nanoparticles. Appl. Surf. Sci. 2010, 256, 3093–3097. [Google Scholar] [CrossRef]
  46. Husain, S.; Irfansyah, M.; Haryanti, N.H.; Suryajaya, S.; Arjo, S.; Maddu, A. Synthesis and characterization of Fe3O4 magnetic nanoparticles from iron ore. J. Phys. Conf. Ser. 2019, 1242, 012021. [Google Scholar] [CrossRef]
  47. Li, W. Facile synthesis of monodisperse Bi2O3 nanoparticles. Mater. Chem. Phys. 2006, 99, 174–180. [Google Scholar] [CrossRef]
  48. Labib, S. Preparation, characterization and photocatalytic properties of doped and undoped Bi2O3. J. Saudi Chem. Soc. 2017, 21, 664–672. [Google Scholar] [CrossRef]
  49. Wu, X.; Wang, X.; Chen, X.; Yang, X.; Ma, Q.; Xu, G.; Yu, L.; Ding, J. Injectable and thermosensitive hydrogels mediating a universal macromolecular contrast agent with radiopacity for noninvasive imaging of deep tissues. Bioact. Mater. 2021, 6, 4717–4728. [Google Scholar] [CrossRef] [PubMed]
  50. Dukic, W.; Delija, B.; Derossi, D.; Dadic, I. Radiopacity of composite dental materials using a digital X-ray system. Dent. Mater. J. 2012, 31, 47–53. [Google Scholar] [CrossRef] [PubMed]
  51. Hitij, T.; Fidler, A. Radiopacity of dental restorative materials. Clin. Oral Investig. 2013, 17, 1167–1177. [Google Scholar] [CrossRef] [PubMed]
  52. Fu, N.; Li, A.; Zhang, J.; Zhang, P.; Zhang, H.; Yang, S.; Zhang, J. Liposome-camouflaged iodinated mesoporous silica nanoparticles with high loading capacity, high hemodynamic stability, high biocompatibility and high radiopacity. Int. J. Pharmaceut. 2024, 650, 123700. [Google Scholar] [CrossRef] [PubMed]
  53. Emonde, C.V.; Eggers, M.E.; Wichmann, M.; Hurschler, C.; Ettinger, M.; Denkena, B. Radiopacity Enhancements in Polymeric Implant Biomaterials: A Comprehensive Literature Review. ACS Biomater. Sci. Eng. 2024, 10, 1323–1334. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Solid-state reaction synthesis of Bi25FeO40.
Figure 1. Solid-state reaction synthesis of Bi25FeO40.
Crystals 14 00706 g001
Figure 2. XRD patterns of Bi25FeO40 synthesis products obtained at 700 °C, 750 °C, and 800 °C. Vertical lines represent the standard XRD pattern of Bi25FeO40.
Figure 2. XRD patterns of Bi25FeO40 synthesis products obtained at 700 °C, 750 °C, and 800 °C. Vertical lines represent the standard XRD pattern of Bi25FeO40.
Crystals 14 00706 g002
Figure 3. FTIR spectra of Bi25FeO40 synthesis products obtained at 700 °C, 750 °C, and 800 °C.
Figure 3. FTIR spectra of Bi25FeO40 synthesis products obtained at 700 °C, 750 °C, and 800 °C.
Crystals 14 00706 g003
Figure 4. SEM micrographs of Bi25FeO40 synthesis products obtained at different temperatures (a) 550 °C; (b) 700 °C; (c) 750 °C; (d) 850 °C.
Figure 4. SEM micrographs of Bi25FeO40 synthesis products obtained at different temperatures (a) 550 °C; (b) 700 °C; (c) 750 °C; (d) 850 °C.
Crystals 14 00706 g004
Figure 5. Photos of reference samples prepared after 1 h and 12 h showing the stability of prepared reference materials.
Figure 5. Photos of reference samples prepared after 1 h and 12 h showing the stability of prepared reference materials.
Crystals 14 00706 g005
Figure 6. Photos of samples with iron–sillenite prepared after 1 h and 12 h.
Figure 6. Photos of samples with iron–sillenite prepared after 1 h and 12 h.
Crystals 14 00706 g006
Figure 7. Radiographs of iron–sillenite Bi25FeO40 specimens mixed with BaSO4 and Laponite.
Figure 7. Radiographs of iron–sillenite Bi25FeO40 specimens mixed with BaSO4 and Laponite.
Crystals 14 00706 g007
Figure 8. Grayscale value dependence on Al mm.
Figure 8. Grayscale value dependence on Al mm.
Crystals 14 00706 g008
Table 1. The reference samples used for the radiopacity measurements.
Table 1. The reference samples used for the radiopacity measurements.
Name of Control Samplem(BaSO4), Gm(Laponite Solution), G
A10.10.9
B10.20.8
C10.30.7
A20.21.8
B20.41.6
C20.61.4
Table 2. The radiopacity of reference samples.
Table 2. The radiopacity of reference samples.
Name of Control SampleGrayscale Valuemm Al
A110.7730.355
B184.5963.387
C1142.9102.922
A253.8930.681
B283.0084.257
C2159.5115.092
Table 3. The radiopacity of iron–sillenite Bi25FeO40 specimens.
Table 3. The radiopacity of iron–sillenite Bi25FeO40 specimens.
Name of Control SampleInvestigated SamplesBackgroundIron–Sillenite
Grayscale Valuemm AlGrayscale Valuemm Almm Al
A368.406 ± 6.5342.27960.143 ± 6.3821.9200.359
B3102.346 ± 7.5593.37975.137 ± 6.2422.2741.105
C3100.070 ± 7.6543.79060.731 ± 6.6521.9671.823
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

Vistorskaja, D.; Yang, J.-C.; Wu, Y.-T.; Chang, L.-Y.; Lu, P.-W.; Zarkov, A.; Grigoraviciute, I.; Kareiva, A. Synthesis and Characterization of Iron–Sillenite for Application as an XRD/MRI Dual-Contrast Agent. Crystals 2024, 14, 706. https://doi.org/10.3390/cryst14080706

AMA Style

Vistorskaja D, Yang J-C, Wu Y-T, Chang L-Y, Lu P-W, Zarkov A, Grigoraviciute I, Kareiva A. Synthesis and Characterization of Iron–Sillenite for Application as an XRD/MRI Dual-Contrast Agent. Crystals. 2024; 14(8):706. https://doi.org/10.3390/cryst14080706

Chicago/Turabian Style

Vistorskaja, Diana, Jen-Chang Yang, Yu-Tzu Wu, Liang-Yu Chang, Po-Wen Lu, Aleksej Zarkov, Inga Grigoraviciute, and Aivaras Kareiva. 2024. "Synthesis and Characterization of Iron–Sillenite for Application as an XRD/MRI Dual-Contrast Agent" Crystals 14, no. 8: 706. https://doi.org/10.3390/cryst14080706

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