Next Article in Journal
Synergistic Effect of Nanofluids and Surfactants on Heavy Oil Recovery and Oil-Wet Calcite Wettability
Previous Article in Journal
Effect of the Processing on the Resistance–Strain Response of Multiwalled Carbon Nanotube/Natural Rubber Composites for Use in Large Deformation Sensors
Previous Article in Special Issue
Quenching of the Eu3+ Luminescence by Cu2+ Ions in the Nanosized Hydroxyapatite Designed for Future Bio-Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Editorial

Nanostructural Materials with Rare Earth Ions: Synthesis, Physicochemical Characterization, Modification and Applications

Institute of Low Temperature and Structure Research, PAS, Okolna 2, 50-422 Wroclaw, Poland
Nanomaterials 2021, 11(7), 1848; https://doi.org/10.3390/nano11071848
Submission received: 7 July 2021 / Revised: 9 July 2021 / Accepted: 13 July 2021 / Published: 16 July 2021
The success of nanotechnology in the field of physical, chemical and medical sciences has started revolutionizing the drug delivery science and theranostics (therapy and diagnostics) [1,2]. The specific advantages include superior pharmacodynamics, pharmacokinetics, reduced toxicity and improved targeting capability. This approach has great potential to produce novel diagnostics and therapeutics—theranostic—because of nanomaterials show unexpected and interesting chemical and physical properties different from those of the original in the micro-sized scale [3]. Therapies combining the use of bioactive materials and progenitor cells or an active substance become clinical reality, increasing the prospects for the development of engineering and regenerative medicine [4]. One of these perspectives is a diagnostics and personalized therapy, i.e., theranostics [5].
In this case of the active substance, the drug-delivery vehicle, as a critical quality attribute in the drug delivery science, needs special attention for the formulation development, which can be successfully achieved via nanotechnology. Drugs incorporated in nanocarriers, either physically entrapped or chemically tethered, have the potential to target the physiological zone of the disorder sparing normal cells from collateral consequences. Targeting several molecular mechanisms, for either treatment or prevention of difficult-to-treat diseases, for the design of various nanotechnology-based drug delivery systems is one of the prime focuses of the formulation scientist at the present juncture.
Much attention has been devoted to developing new drug-delivery systems with many advantages compared with the conventional forms of dosage, such as, among others, enhanced bioavailability, greater efficiency, lower toxicity, controlled release [6,7,8,9,10,11,12]. An ideal drug-delivery system should be characterized by: (1) maximum biocompatibility and minimal antigenic properties [13]; (2) appropriate particle size, which is important for the particles to reach a particular location in the body due to the size of the vessels of the human circulatory system [14]; (3) the ability to transport the desired drug molecules to the targeted cells or tissues and release them in a controlled manner [15]. So far, different types of drug-delivery systems have been developed, such as, i.a. biodegradable polymers [16], xerogels [17], hydrogels [18], mesoporous materials [4,11]. Among different drug-delivery systems, mesoporous materials (such as SBA-15, MCM-41 and mesoporous silica nanoparticles) have gained increasing interest, particularly as drug storage and release hosts due to their unique surface and textural properties [14,19,20].
Materials designed for biomedical applications should be characterized by a high sensitivity and specificity, a lack of functional interference with the sample, photochemical stability, non-toxicity, long time of storage and, as far as possible, detection of a substance in the presence of others. Moreover, nanoscale materials have been exploited as active components in a wide range of technological applications in the biomedical field [21,22,23,24,25]. Particularly, in the field of biomedicine, nanoparticles can be used as drug-delivery vehicles that can target tissues or cells [13,24] and can be functionalized with special characteristics (such as magnetization, fluorescence and near-infrared absorption) for qualitative or quantitative detection of tumor cells [23,25,26,27].
It is well-known that nanoscale fluorescent materials have attracted much interest due to the increasing demand for efficient photosensitive materials not only for sophisticated optoelectronic and photonic devices but also for a broad range of biomedical applications [28,29,30,31,32,33,34]. In biomedical areas, luminescent materials, mainly including fluorescent organic molecules [35,36] and semiconductor nanoparticles [37,38], have been widely investigated in biological staining and diagnostics. However, some serious problems of photobleaching and quenching of fluorescent organic molecules and the toxicity of semiconductor quantum dots are critically evident and have seriously limited their applications in biomedical areas [38,39]. Furthermore, high performance in function-specific biological applications requires that the composites possess some unique characteristics, such as uniform morphology, large surface areas, good dispersion, etc. [39]. Recently, a class of stable, efficient and self-activated luminescent materials whose emission is induced by the defects or impurities in host lattices, has been prepared by various synthesis routes [40,41,42,43]. These novel self-activated inorganic materials may be a promising fluorescent material for biodetection due to their good optical properties and nontoxicity.
Apatites are inorganic compounds with a general formula M10(XO4)6Y2, where M represents divalent cations (e.g., Ca2+, Sr2+, etc.), XO4 = PO43−, VO43−, etc. and Y represents anions: F, OH, Cl, Br, etc. The hexagonal structure in apatites belongs to P63/m space group and allows the cations to localize in the 4(f) and 6(h) positions [44] and is able to accommodate a variety of univalent cations as substituents. In that case, charge compensation, proposed by P. Martin and et al. [45], allows explaining the substitution of divalent calcium ions to trivalent lanthanide ions in apatite with a simple mechanism. It is worth mentioning that apatites themselves, such as calcium apatites Ca10(PO4)6(Y)2, are biocompatible and are natural building blocks for bones and teeth [46]. This feature combined with highly photostable luminescent properties of rare-earth dopants, makes nanocrystalline apatites highly attractive as luminescent bio-labels [47]. However, these materials have not been extensively synthesized or examined in the nanocrystalline form [48] which is a prerequisite for being internalized by cells for bio-imaging or sensing applications [49].
Several strategies have been developed in the synthesis of nanoparticles so far, involving such techniques as microemulsion, precipitation, thermal decomposition, chemical vapor deposition and others. However, the best results and control over particle size, crystallinity and purity can be ensured using microwave technology [50]. For instance, our group was able to obtain highly crystalline, phase pure, bio-compatible uniform and low agglomerated nano-apatites such as Ca10(PO4)6(OH)2 for bio-applications [51,52]. Another important feature is that the materials were produced in environmentally friendly conditions in ethylene glycol solution that is non-toxic for living organisms. Thus, this strategy seems to be very attractive for the synthesis of luminescent or multifunctional materials offering the possibility of bio-imaging measurement. The proposed synthesis technique allows for thorough control over the desired composition (as it was shown in the article [53]) which cannot be simply achieved using other techniques. Moreover, the proposed compounds can be considered as non-toxic due to their insolubility in body fluids and high chemical stability. It is well known that the solubility of oxide nanoparticles is one of the most important factors of their toxicity related to their chemical composition [54]. For example, the toxic effect of iron oxide nanoparticles originates mainly from the catalytic production of free radicals through Fenton type reaction [55]. To date, quantum dots (QD) characterized by high absorbance, high quantum yield, narrow emission bands and high resistance to photobleaching were considered as the most promising materials for FI applications in medicine. Currently, the main issue regarding QDs and their biomedical applications is their extreme toxicity (semiconductors—derivatives of highly toxic heavy metals such as Cd or Pb) [56]. One of the promising alternatives is offered by the application of inorganic compounds such as apatites doped or co-doped with optically active rare earth metals for bio-imaging [57].
Furthermore, calcium is the fifth most abundant element by mass in the human body (1.4–1.66%) where it is a common cellular ionic messenger with many functions and serves also as a structural element in bones (hydroxyapatites—99%) [58]. Calcium and its compounds play an important role in controlling numerous biological processes in living systems. Concentrations of free Ca2+ in biological cells are widely studied with fluorescent probes. The probes have a high selectivity for free calcium and exhibit marked changes in their photophysical properties upon binding. In particular, changes in fluorescence intensity (intensity probes) or spectral shift (ratio probes) upon binding to Ca2+ are monitored. The main drawback of intensity probes is that the intensity of fluorescence is affected by both the probe concentration and the free Ca2+ concentration. Consequently, a quantitative determination of Ca2+ distributions requires the probes to be distributed homogeneously in the sample. Conventional quantitative determinations of Ca2+ concentration with ratio probes overcomes the dependence on local probe concentration by exploiting ratiometric procedures using excitation or detection at two wavelengths [59,60]. The advent of fluorescence lifetime imaging techniques [1,61,62,63,64] opens new horizons for the quantitative determination for bio-imaging, in particular using intensity probes [65]. Fluorescence lifetime imaging is determined by factors such as the chemical environment of a fluorescent molecule and thus provides valuable information about its ion binding states. Importantly, since the lifetime is independent of fluorescence intensity, such measurements have wide-ranging applications to samples in which the probes have an inhomogeneous distribution. An additional advantage of the lifetime imaging technique is that the images are not compromised by photobleaching and absorption effects.
Nanocrystalline probes doped with lanthanide ions based on apatites meet these requirements [66]. Their narrow emission lines as well as long life-times render them suitable for use as luminescent markers in biology and medicine [67]. Therefore, this strategy seems to be very attractive for the complete elimination of the effects associated with local concentration of ions in the sample. Moreover, the surface functionalization of nanomaterials with biologically active organic ligands results in a better stability of the colloidal dispersion. It will contribute to measurable progress in the possible extension of bio-imaging techniques. Independently of the scientific goal related to theranostics, the synthesis and study of spectroscopic properties of lanthanide-ion doped apatites could also be an important area of research.

Funding

The research was funded by the National Science Centre (NCN) within the projects “Preparation and characterisation of biocomposites based on nanoapatites for theranostics” (No. UMO-2015/19/B/ST5/01330).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Rafal J. Wiglusz would like to thank all authors who submitted their research to this Special Issue, the referees who reviewed the submitted manuscripts and the Assistant Editor, Tina Tian, who made it all work.

Conflicts of Interest

Author declares no conflict of interest.

References

  1. Szyszka, K.; Targońska, S.; Lewińska, A.; Watras, A.; Wiglusz, R.J. Quenching of the Eu3+ luminescence by Cu2+ ions in the nanosized hydroxyapatite designed for future bio-detection. Nanomaterials 2021, 11, 464. [Google Scholar] [CrossRef]
  2. Zienkiewicz, J.A.; Strzep, A.; Jedrzkiewicz, D.; Nowak, N.; Rewak-Soroczynska, J.; Watras, A.; Ejfler, J.; Wiglusz, R.J. Preparation and Characterization of Self-Assembled Poly(l-Lactide) on the Surface of β-Tricalcium Diphosphate(V) for Bone Tissue Theranostics. Nanomaterials 2020, 10, 331. [Google Scholar] [CrossRef] [Green Version]
  3. Saiz, E.; Zimmermann, E.A.; Lee, J.S.; Wegst, U.G.K.; Tomsia, A.P. Perspectives on the role of nanotechnology in bone tissue engineering. Dent. Mater. 2013, 29, 103–115. [Google Scholar] [CrossRef] [Green Version]
  4. Zakrzewski, W.; Dobrzynski, M.; Rybak, Z.; Szymonowicz, M.; Wiglusz, R.J. Selected Nanomaterials’ Application Enhanced with the Use of Stem Cells in Acceleration of Alveolar Bone Regeneration during Augmentation Process. Nanomaterials 2020, 10, 1216. [Google Scholar] [CrossRef] [PubMed]
  5. Watras, A.; Wujczyk, M.; Roecken, M.; Kucharczyk, K.; Marycz, K.; Wiglusz, R.J. Investigation of Pyrophosphates KYP2O7 Co-Doped with Lanthanide Ions Useful for Theranostics. Nanomaterials 2019, 9, 1597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Fischer, K.E.; Alemán, B.J.; Tao, S.L.; Daniels, R.H.; Li, E.M.; Bünger, M.D.; Nagaraj, G.; Singh, P.; Zettl, A.; Desai, T.A. Biomimetic Nanowire Coatings for Next Generation Adhesive Drug Delivery Systems. Nano Lett. 2009, 9, 716–720. [Google Scholar] [CrossRef] [Green Version]
  7. Sobierajska, P.; Serwotka-Suszczak, A.; Szymanski, D.; Marycz, K.; Wiglusz, R.J. Nanohydroxyapatite-Mediated Imatinib Delivery for Specific Anticancer Applications. Molecules 2020, 25, 4602. [Google Scholar] [CrossRef]
  8. Vivero-Escoto, J.L.; Slowing, I.I.; Wu, C.W.; Lin, V.S.Y. Photoinduced intracellular controlled release drug delivery in human cells by gold-capped mesoporous silica nanosphere. J. Am. Chem. Soc. 2009, 131, 3462–3463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Wei, W.; Ma, G.H.; Hu, G.; Yu, D.; Mcleish, T.; Su, Z.G.; Shen, Z.Y. Preparation of hierarchical hollow CaCo3 particles and the application as anticancer drug carrier. J. Am. Chem. Soc. 2008, 130, 15808–15810. [Google Scholar] [CrossRef] [PubMed]
  10. Yang, Q.; Wang, S.; Fan, P.; Wang, L.; Di, Y.; Lin, K.; Xiao, F.S. pH-responsive carrier system based on carboxylic acid modified mesoporous silica and polyelectrolyte for drug delivery. Chem. Mater. 2005, 17, 5999–6003. [Google Scholar] [CrossRef]
  11. Zhao, W.; Chen, H.; Li, Y.; Li, A.; Lang, M.; Shi, J. Uniform rattle-type hollow magnetic mesoporous spheres as drug delivery carriers and their sustained-release property. Adv. Funct. Mater. 2008, 18, 2780–2788. [Google Scholar] [CrossRef]
  12. Rösler, A.; Vandermeulen, G.W.M.; Klok, H.A. Advanced drug delivery devices via self-assembly of amphiphilic block copolymers. Adv. Drug Deliv. Rev. 2001, 53, 95–108. [Google Scholar] [CrossRef]
  13. Chen, F.H.; Gao, Q.; Ni, J.Z. The grafting and release behavior of doxorubincin from Fe3O4@SiO2 core-shell structure nanoparticles via an acid cleaving amide bond: The potential for magnetic targeting drug delivery. Nanotechnology 2008, 19. [Google Scholar] [CrossRef]
  14. Ritter, J.A.; Ebner, A.D.; Daniel, K.D.; Stewart, K.L. Application of high gradient magnetic separation principles to magnetic drug targeting. J. Magn. Magn. Mater. 2004, 280, 184–201. [Google Scholar] [CrossRef]
  15. Sobierajska, P.; Dorotkiewicz-Jach, A.; Zawisza, K.; Okal, J.; Olszak, T.; Drulis-Kawa, Z.; Wiglusz, R.J. Preparation and antimicrobial activity of the porous hydroxyapatite nanoceramics. J. Alloys Compd. 2018, 748, 179–187. [Google Scholar] [CrossRef]
  16. Uhrich, K.E.; Cannizzaro, S.M.; Langer, R.S.; Shakesheff, K.M. Polymeric Systems for Controlled Drug Release. Chem. Rev. 1999, 99, 3181–3198. [Google Scholar] [CrossRef] [PubMed]
  17. Yang, H.H.; Zhu, Q.Z.; Qu, H.Y.; Chen, X.L.; Ding, M.T.; Xu, J.G. Flow injection fluorescence immunoassay for gentamicin using sol-gel-derived mesoporous biomaterial. Anal. Biochem. 2002, 308, 71–76. [Google Scholar] [CrossRef]
  18. Caliceti, P.; Salmaso, S.; Lante, A.; Yoshida, M.; Katakai, R.; Martellini, F.; Mei, L.H.I.; Carenza, M. Controlled release of biomolecules from temperature-sensitive hydrogels prepared by radiation polymerization. J. Control. Release 2001, 75, 173–181. [Google Scholar] [CrossRef]
  19. Yang, P.; Huang, S.; Kong, D.; Lin, J.; Fu, H. Luminescence functionalization of SBA-15 by YVO4:Eu3+ as a novel drug delivery system. Inorg. Chem. 2007, 46, 3203–3211. [Google Scholar] [CrossRef]
  20. Vallet-Regi, M.; Rámila, A.; Del Real, R.P.; Pérez-Pariente, J. A new property of MCM-41: Drug delivery system. Chem. Mater. 2001, 13, 308–311. [Google Scholar] [CrossRef]
  21. Slowing, I.I.; Trewyn, B.G.; Giri, S.; Lin, V.S.Y. Mesoporous silica nanoparticles for drug delivery and biosensing applications. Adv. Funct. Mater. 2007, 17, 1225–1236. [Google Scholar] [CrossRef]
  22. Manna, L.; Milliron, D.J.; Meisel, A.; Scher, E.C.; Alivisatos, A.P. Controlled growth of tetrapod-branched inorganic nanocrystals. Nat. Mater. 2003, 2, 382–385. [Google Scholar] [CrossRef]
  23. Shi, D. Integrated multifunctional nanosystems for medical diagnosis and treatment. Adv. Funct. Mater. 2009, 19, 3356–3373. [Google Scholar] [CrossRef]
  24. Farokhzad, O.C.; Langer, R. Impact of nanotechnology on drug delivery. ACS Nano 2009, 3, 16–20. [Google Scholar] [CrossRef] [PubMed]
  25. Walt, D.R. Miniature analytical methods for medical diagnostics. Science 2005, 308, 217–219. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, W.; Shi, D.; Lian, J.; Guo, Y.; Liu, G.; Wang, L.; Ewing, R.C. Luminescent hydroxylapatite nanoparticles by surface functionalization. Appl. Phys. Lett. 2006, 89, 183106. [Google Scholar] [CrossRef] [Green Version]
  27. Burns, A.; Ow, H.; Wiesner, U. Fluorescent core–shell silica nanoparticles: Towards “Lab on a Particle” architectures for nanobiotechnology. Chem. Soc. Rev. 2006, 35, 1028–1042. [Google Scholar] [CrossRef] [PubMed]
  28. Diamente, P.R.; Burke, R.D.; Van Veggel, F.C.J.M. Bioconjugation of Ln3+-Doped LaF3 nanoparticles to avidin. Langmuir 2006, 22, 1782–1788. [Google Scholar] [CrossRef]
  29. Dong, C.; Van Veggel, F.C.J.M. Cation exchange in lanthanide fluoride nanoparticles. ACS Nano 2009, 3, 123–130. [Google Scholar] [CrossRef]
  30. Syamchand, S.S.; Sony, G. Europium enabled luminescent nanoparticles for biomedical applications. J. Lumin. 2015, 165, 190–215. [Google Scholar] [CrossRef]
  31. Jeong, J.; Cho, M.; Lim, Y.T.; Song, N.W.; Chung, B.H. Synthesis and characterization of a photoluminescent nanoparticle based on fullerene-silica hybridization. Angew. Chem. Int. Ed. 2009, 48, 5296–5299. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, M.; Mi, C.C.; Wang, W.X.; Liu, C.H.; Wu, Y.F.; Xu, Z.R.; Mao, C.B.; Xu, S.K. Immunolabeling and NIR-excited fluorescent imaging of HeLa cells by using NaYF 4:Yb,Er upconversion nanoparticles. ACS Nano 2009, 3, 1580–1586. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, F.; Tan, W.B.; Zhang, Y.; Fan, X.; Wang, M. Luminescent nanomaterials for biological labelling. Nanotechnology 2006, 17, 1–13. [Google Scholar] [CrossRef]
  34. Ninjbadgar, T.; Garnweitner, G.; Boärger, A.; Goldenberg, L.M.; Sakhno, O.V.; Stumpe, J. Synthesis of luminescent ZrO2:Eu3+ nanoparticles and their holographic sub-micrometer patterning in polymer composites. Adv. Funct. Mater. 2009, 19, 1819–1825. [Google Scholar] [CrossRef]
  35. Kim, J.; Kim, H.S.; Lee, N.; Kim, T.; Kim, H.; Yu, T.; Song, I.C.; Moon, W.K.; Hyeon, T. Multifunctional Uniform Nanoparticles Composed of a Magnetite Nanocrystal Core and a Mesoporous Silica Shell for Magnetic Resonance and Fluorescence Imaging and for Drug Delivery. Angew. Chemie 2008, 120, 8566–8569. [Google Scholar] [CrossRef]
  36. Zumbuehl, A.; Jeannerat, D.; Martin, S.E.; Sohrmann, M.; Stano, P.; Vigassy, T.; Clark, D.D.; Hussey, S.L.; Peter, M.; Peterson, B.R.; et al. An amphotericin B-fluorescein conjugate as a powerful probe for biochemical studies of the membrane. Angew. Chemie Int. Ed. 2004, 43, 5181–5185. [Google Scholar] [CrossRef]
  37. Gao, X.; Nie, S. Molecular profiling of single cells and tissue specimens with quantum dots. Trends Biotechnol. 2003, 21, 371–373. [Google Scholar] [CrossRef]
  38. Yong, K.T.; Ding, H.; Roy, I.; Law, W.C.; Bergey, E.J.; Maitra, A.; Prasad, P.N. Imaging pancreatic cancer using bioconjugated inp quantum dots. ACS Nano 2009, 3, 502–510. [Google Scholar] [CrossRef]
  39. Yang, P.; Quan, Z.; Hou, Z.; Li, C.; Kang, X.; Cheng, Z.; Lin, J. A magnetic, luminescent and mesoporous core-shell structured composite material as drug carrier. Biomaterials 2009, 30, 4786–4795. [Google Scholar] [CrossRef]
  40. Green, W.H.; Le, K.P.; Grey, J.; Au, T.T.; Sailor, M.J. White phosphors from a silicate-carboxylate sol-gel precursor that lack metal activator ions. Science 1997, 276, 1826–1828. [Google Scholar] [CrossRef]
  41. Zawisza, K.; Strzep, A.; Wiglusz, R.J. Influence of annealing temperature on the spectroscopic properties of hydroxyapatite analogues doped with Eu3+. New J. Chem. 2017, 41, 9990–9999. [Google Scholar] [CrossRef]
  42. Ogi, T.; Kaihatsu, Y.; Iskandar, F.; Wang, W.N.; Okuyama, K. Facile synthesis of new full-color-emitting BCNO phosphors with high quantum efficiency. Adv. Mater. 2008, 20, 3235–3238. [Google Scholar] [CrossRef]
  43. Angelov, S.; Stoyanova, R.; Dafinova, R.; Kabasanov, K. Luminescence and EPR studies on strontium carbonate obtained by thermal decomposition of strontium oxalate. J. Phys. Chem. Solids 1986, 47, 409–412. [Google Scholar] [CrossRef]
  44. Sudarsanan, K.; Young, R.A. Significant precision in crystal structural details. Holly Springs hydroxyapatite. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1969, 25, 1534–1543. [Google Scholar] [CrossRef]
  45. Martin, P.; Carlot, G.; Chevarier, A.; Den-Auwer, C.; Panczer, G. Mechanisms involved in thermal diffusion of rare earth elements in apatite. J. Nucl. Mater. 1999, 275, 268–276. [Google Scholar] [CrossRef] [Green Version]
  46. Zhang, H.; Ye, X.J.; Li, J.S. Preparation and biocompatibility evaluation of apatite/wollastonite-derived porous bioactive glass ceramic scaffolds. Biomed. Mater. 2009, 4, 045007. [Google Scholar] [CrossRef] [PubMed]
  47. Doat, A.; Pellé, F.; Gardant, N.; Lebugle, A. Synthesis of luminescent bioapatite nanoparticles for utilization as a biological probe. J. Solid State Chem. 2004, 177, 1179–1187. [Google Scholar] [CrossRef]
  48. Lebugle, A.; Pellé, F.; Charvillat, C.; Rousselot, I.; Chane-Ching, J.Y. Colloidal and monocrystalline Ln3+ doped apatite calcium phosphate as biocompatible fluorescent probes. Chem. Commun. 2006, 6, 606–608. [Google Scholar] [CrossRef] [Green Version]
  49. Lubojanski, A.; Dobrzynski, M.; Nowak, N.; Rewak-Soroczynska, J.; Sztyler, K.; Zakrzewski, W.; Dobrzynski, W.; Szymonowicz, M.; Rybak, Z.; Wiglusz, K.; et al. Application of Selected Nanomaterials and Ozone in Modern Clinical Dentistry. Nanomaterials 2021, 11, 259. [Google Scholar] [CrossRef]
  50. Wiglusz, R.J.; Grzyb, T.; Lis, S.; Strek, W. Hydrothermal preparation and photoluminescent properties of MgAl2O4:Eu3+ spinel nanocrystals. J. Lumin. 2010, 130, 434–441. [Google Scholar] [CrossRef]
  51. Wiglusz, R.J.; Kedziora, A.; Lukowiak, A.; Doroszkiewicz, W.; Strek, W. Hydroxyapatites and Europium(III) doped hydroxyapatites as a carrier of silver nanoparticles and their antimicrobial activity. J. Biomed. Nanotechnol. 2012, 8, 605–612. [Google Scholar] [CrossRef] [PubMed]
  52. Targonska, S.; Rewak-Soroczynska, J.; Piecuch, A.; Paluch, E.; Szymanski, D.; Wiglusz, R.J. Preparation of a New Biocomposite Designed for Cartilage Grafting with Antibiofilm Activity. ACS Omega 2020, 5, 24546–24557. [Google Scholar] [CrossRef] [PubMed]
  53. Pazik, R.; Wiglusz, R.J.; Strek, W. Luminescence properties of BaTiO3:Eu3+ obtained via microwave stimulated hydrothermal method. Mater. Res. Bull. 2009, 44, 1328–1333. [Google Scholar] [CrossRef]
  54. Brunner, T.J.; Wick, P.; Manser, P.; Spohn, P.; Grass, R.N.; Limbach, L.K.; Bruinink, A.; Stark, W.J. In vitro cytotoxicity of oxide nanoparticles: Comparison to asbestos, silica, and the effect of particle solubility. Environ. Sci. Technol. 2006, 40, 4374–4381. [Google Scholar] [CrossRef] [PubMed]
  55. Nũnez, M.T.; Garate, M.A.; Arredondo, M.; Tapia, V.; Muñoz, P. The cellular mechanisms of body iron homeostasis. Biol. Res. 2000, 33, 133–142. [Google Scholar] [CrossRef]
  56. Sousanis, A.; Poelman, D.; Smet, P.F. SmS/EuS/SmS Tri-Layer Thin Films: The Role of Diffusion in the Pressure Triggered Semiconductor-Metal Transition. Nanomaterials 2019, 9, 1513. [Google Scholar] [CrossRef] [Green Version]
  57. Cao, H.; Zhang, L.; Zheng, H.; Wang, Z. Hydroxyapatite nanocrystals for biomedical applications. J. Phys. Chem. C 2010, 114, 18352–18357. [Google Scholar] [CrossRef]
  58. Tordoff, M.G. Calcium: Taste, intake, and appetite. Physiol. Rev. 2001, 81, 1567–1597. [Google Scholar] [CrossRef]
  59. Grynkiewicz, G.; Poenie, M.; Tsien, R.Y. A New Generation of Ca2+ Indicators with Greatly Improved Fluorescence Properties. J. Biol. Chem. 1985, 260, 3440–3450. [Google Scholar] [CrossRef]
  60. Tsien, R.Y.; Poenie, M. Fluorescence ratio imaging: A new window into intracellular ionic signaling. Trends Biochem. Sci. 1986, 11, 450–455. [Google Scholar] [CrossRef]
  61. Buurman, E.P.; Sanders, R.; Draaijer, A.; Gerritsen, H.C.; van Veen, J.J.F.; Houpt, P.M.; Levine, Y.K. Fluorescence lifetime imaging using a confocal laser scanning microscope. Scanning 1992, 14, 155–159. [Google Scholar] [CrossRef]
  62. Lakowicz, J.R.; Szmacinski, H.; Nowaczyk, K.; Johnson, M.L. Fluorescence lifetime imaging of free and protein-bound NADH. Proc. Natl. Acad. Sci. USA 1992, 89, 1271–1275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Morgan, C.G.; Mitchell, A.C.; Murray, J.G. Fluorescence decay time imaging using an imaging photon detector with a radio frequency photon correlation system. In Proceedings of the Time-Resolved Laser Spectroscopy in Biochemistry II; SPIE: Bellingham, WA, USA, 1990; Volume 1204, pp. 798–807. [Google Scholar] [CrossRef]
  64. Wiatrak, B.; Sobierajska, P.; Szandruk-Bender, M.; Jawien, P.; Janeczek, M.; Dobrzynski, M.; Pistor, P.; Szelag, A.; Wiglusz, R.J. Nanohydroxyapatite as a biomaterial for peripheral nerve regeneration after mechanical damage—in vitro study. Int. J. Mol. Sci. 2021, 22, 4454. [Google Scholar] [CrossRef]
  65. Zakrzewski, W.; Dobrzynski, M.; Nowicka, J.; Pajaczkowska, M.; Szymonowicz, M.; Targonska, S.; Sobierajska, P.; Wiglusz, K.; Dobrzynski, W.; Lubojanski, A.; et al. The influence of ozonated olive oil-loaded and copper-doped nanohydroxyapatites on planktonic forms of microorganisms. Nanomaterials 2020, 10, 1997. [Google Scholar] [CrossRef]
  66. Wiglusz, R.J.; Bednarkiewicz, A.; Strek, W. Synthesis and optical properties of Eu3+ ion doped nanocrystalline hydroxya patites embedded in PMMA matrix. J. Rare Earths 2011, 29, 1111–1116. [Google Scholar] [CrossRef]
  67. Pazik, R.; Tekoriute, R.; Håkansson, S.; Wiglusz, R.; Strek, W.; Seisenbaeva, G.A.; Gun’ko, Y.K.; Kessler, V.G. Precursor and solvent effects in the nonhydrolytic synthesis of complex oxide nanoparticles for bioimaging applications by the ether elimination (Bradley) reaction. Chem. A Eur. J. 2009, 15, 6820–6826. [Google Scholar] [CrossRef] [PubMed]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wiglusz, R.J. Nanostructural Materials with Rare Earth Ions: Synthesis, Physicochemical Characterization, Modification and Applications. Nanomaterials 2021, 11, 1848. https://doi.org/10.3390/nano11071848

AMA Style

Wiglusz RJ. Nanostructural Materials with Rare Earth Ions: Synthesis, Physicochemical Characterization, Modification and Applications. Nanomaterials. 2021; 11(7):1848. https://doi.org/10.3390/nano11071848

Chicago/Turabian Style

Wiglusz, Rafal J. 2021. "Nanostructural Materials with Rare Earth Ions: Synthesis, Physicochemical Characterization, Modification and Applications" Nanomaterials 11, no. 7: 1848. https://doi.org/10.3390/nano11071848

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