Next Article in Journal
Single-Frame Vignetting Correction for Post-Stitched-Tile Imaging Using VISTAmap
Previous Article in Journal
Unlocking the Potential of Mg-Doped Rare Earth Manganites: Machine Learning-Guided Synthesis and Insights into Structural and Optical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pr3+ Visible to Ultraviolet Upconversion for Antimicrobial Applications

1
National Institute of Research and Development for Electrochemistry and Condensed Matter, Str. Dr. A. Păunescu Podeanu nr.144, 300569 Timisoara, Romania
2
Centre of Excellence for Photoconversion, Vinča Institute of Nuclear Sciences—National Institute of the Republic of Serbia, University of Belgrade, Mike Petrovi12-14, 11000 Belgrade, Serbia
3
Inorganic Photoactive Materials, Institute of Inorganic and Structural Chemistry, Heinrich Heine University Düsseldorf, Universitätsstraße 1, 40225 Düsseldorf, Germany
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(7), 562; https://doi.org/10.3390/nano15070562
Submission received: 17 March 2025 / Revised: 3 April 2025 / Accepted: 5 April 2025 / Published: 6 April 2025
(This article belongs to the Section Nanophotonics Materials and Devices)

Abstract

:
This paper addresses the upconversion of blue light to ultraviolet-C (UVC) with Pr3+-activated materials for antibacterial applications of UVC. It discusses the processes through which UV radiation provides biocidal effects on microorganisms, along with the most popular UVC sources employed in these processes. We describe the electronic and optical properties of the Pr3+ ion, emphasizing the conditions the host material must meet to obtain broad and intense emission in the UVC from parity-allowed transitions from the 4f5d levels and provide a list of materials that fulfill these conditions. This paper also delineates lanthanide-based upconversion, focusing on Pr3+ blue to UVC upconversion via the 3P0 and 1D2 intermediate states, and suggests routes for improving the quantum efficiency of the process. We review literature related to the use of upconversion materials in antimicrobial photodynamic treatments and for the blue to UVC upconversion germicidal effects. Further, we propose the spectral overlap between the UVC emission of Pr3+ materials and the germicidal effectiveness curve as a criterion for assessing the potential of these materials in antimicrobial applications. Finally, this paper briefly assesses the toxicity of materials commonly used in the preparation of upconversion materials.

Graphical Abstract

1. Introduction

Recent years have seen increasing adoption of non-thermal and chemical-free disinfection methods, such as ultraviolet (UV)–C light, pulsed UV, cold plasma, and ultrasonic waves. UV radiation effectively combats pathogenic microorganisms by directly interacting with their deoxyribonucleic acid (DNA), leaving no toxic or environmentally stable chemical residues. Since 1877, when the initial scientific report on the germicidal effects of UV radiation was published [1], UV irradiation has been used as an effective antimicrobial strategy in different laboratory and hospital devices, such as UV sterilization lamps and microbiology cabins [2]. It has also been proposed for the treatment of microbial infections in patients [3,4]. Although some microbial species are partially tolerant to UV radiation [3], primarily due to the solar UV levels present in an organism’s natural habitat, it is widely acknowledged that UV radiation adversely affects them [5]. UV radiation can be separated into five wavelength ranges: vacuum-UV (40–190 nm), far-UV (190–220 nm), UVC (220–280 nm), UVB (280–315 nm), and UVA (315–400 nm). The effectiveness of ultraviolet (UV) light in killing germs depends on its wavelength. High-energy radiation (short wavelengths) has a higher potential for damaging microorganisms. However, in 1930, Gates discovered that a characteristic curve of UV radiation bactericidal efficacy has a maximum between 260 and 270 nm [6]. This is a result of specific absorptions of bacterial components that have the largest absorption values around these wavelengths. It should be noted, however, that the wavelength dependence of the UV radiation germicidal effectiveness varies between different species of the same microorganism family, as well as between different types of microorganisms, mostly based on the cell wall structure, DNA arrangement, and repair mechanisms [7], as discussed in more detail in Section 2.
UVC causes the most harm to microorganisms, yet it accounts for only 0.5% of total solar radiation. In fact, ozone in the stratosphere absorbs virtually all the vacuum-UV, far-UV, and UVC; half of the UVB; and some of the UVA radiation before it reaches the Earth’s surface [8]. As a result, we must use artificial UV light sources for disinfecting water, air, and surfaces, as well as for treating tissue infections. The most used sources are discussed in Section 3. A contemporary approach to UV light generation is based on shorter-than-excitation-wavelength (STEW) processes, in which photons emitted by a material have higher energy (shorter wavelengths) than the photons that produced them. In Section 4, we discuss the most widely used STEW process, which is based on lanthanide-facilitated upconversion. Section 5 discusses the application of near-infrared-to-visible lanthanide upconversion nanoparticles in antimicrobial contexts, primarily in antimicrobial photodynamic treatment. This work emphasizes visible-to-UV lanthanide-facilitated upconversion and its germicidal applications (Section 6), primarily focusing on Pr3+ blue-to-UVC upconversion. Several reviews published in recent years [9,10,11] on Pr3+ blue-to-UVC upconversion signify the importance of this fast-growing research field. Here, we cover the electronic and optical properties of Pr3+ ions in Section 7 and explain the mechanisms behind Pr3+ blue-to-UVC upconversion in Section 8, along with the analysis of quantum efficiencies of these processes and strategies aimed at improving the efficiency. Section 9 reviews the germicidal applications of UV radiation produced from Pr3+ upconversion, highlighting several instances from the literature. The final section briefly discusses the toxicity of the materials frequently used for Pr3+ upconversion.

2. On UV Radiation’s Germicidal Effects

Microorganism inactivation is facilitated by different processes depending on the UV light wavelength/energy. UVC is a highly effective bactericidal agent due to its strong absorption by DNA (maximum at 265 nm), where it induces the formation of cyclobutane pyrimidine dimers and 6-4 photoproducts, disrupting replication and transcription processes [12]. These photochemical reactions directly damage DNA, leading to cell death. Buonanno et al. [13] recently demonstrated that long-term UVC (222 nm) exposure can efficiently destroy airborne human coronaviruses. Dai et al. [14] demonstrated that UVC was more effective than nystatin cream in lowering the fungal burden of mouse burns with developed infections. UVB operates through similar photochemical reactions, although it is less efficient than UVC because of the small DNA absorption. However, UVB has a greater penetration depth in tissues than UVC, and it reaches greater depths in water, which is important for water disinfection. When used in combination, UVB and UVC show additive effects rather than synergistic inactivation [15]. UVA has a higher tissue penetration depth and much lower DNA absorption than UVC and UVB. Its principal biocidal activity is the production of reactive oxygen species (ROS), which cause oxidative damage to key biological components such as membranes, proteins, and DNA [16,17,18]. This indirect mechanism results in cellular dysfunction and increased vulnerability to subsequent stressors. A particularly effective strategy involves combining UVA pretreatment with UVC irradiation. During UVA exposure, ROS, especially hydroxyl radicals, are generated within bacterial cells, impairing essential functions such as DNA repair mechanisms. This makes the bacteria more susceptible to UVC-induced DNA damage and inhibits their ability to recover through photoreactivation or dark repair [19,20]. For instance, UVA pretreatment can suppress photoreactivation—a process where UV-damaged DNA is enzymatically repaired under visible light—reducing recovery rates from 60% (with UVC alone) to just 15% [19,20]. Applying a fluence of 17 J/cm2 UVA resulted in a significant enhancement of UVC inactivation, while 52 J/cm2 UVA effectively eliminated the repair shoulder in the fluence–response curve [19]. This enhancement is attributed to oxidative stress caused by UVA, which disables key repair enzymes like photolyase. This combined approach is particularly effective against bacteria like Escherichia coli, as they rely on active metabolic pathways and repair systems. However, viruses such as MS2, which lack such systems, show no additional benefit from UVA pretreatment [19,20].
In highly contaminated environments, bacteria are not isolated on surfaces but exist within bacterial biofilms. These biofilms make the germicidal process more challenging. The penetration depth of UVB and UVA radiation into bacterial biofilms is much greater than that of far-UVC light (200–235 nm) due to the lower absorption coefficient of proteins in this spectral range. This makes UVC light with wavelengths greater than 235 nm dangerous for humans, as it can penetrate the natural layer of dead cells on the skin’s surface and potentially cause genetic mutations with carcinogenic potential. In some cases, UVA radiation can temporarily block cell division, though the bacteria remain viable in the long term. Some pathogens can enter a viable but non-culturable (VBNC) state after UV-C treatment, complicating disinfection assessments [21]. Research highlights the phenomenon of photoreactivation, where some bacteria repair DNA damage when exposed to wavelengths over 330 nm [22]. Bacterial species have different sensitivities to UVC radiation. Research by Cairns [23] shows UV doses from 0.4 to 235 mJ/cm2 effectively deactivate microorganisms, with the extent of DNA damage correlating directly to UV exposure. Martinez-Hernandez et al. [24] found Salmonella enteritidis highly susceptible to UVC, requiring just 2 mJ/cm2 for effective disinfection, compared to higher doses needed for other bacteria. Lim and Harrison [25] confirmed UVC light efficacy (at doses ranging from 0 to 223.1 mJ/cm2) in reducing Salmonella on green tomatoes, demonstrating effectiveness regardless of the tomato’s orientation. Liu et al. [26] noted that water-assisted UV treatments were particularly effective for decontaminating blueberries contaminated with Salmonella, outperforming direct UV exposure methods: 7.9 and 4.6 mW/cm2. Sommer et al. [27] found that E. coli strains required up to 30 mJ/cm2 of UV-C for a 6-log reduction, with strain-specific variability in repair mechanisms. Chun et al. [28] reported that UVC doses between 100 and 800 mJ/cm2 reduced E. coli O157 counts on fresh salads by 2.16 log CFU/g. Allende et al. [29] demonstrated complete bacterial inhibition on fresh products with UV doses between 3 and 8.5 mJ/cm2, though high doses negatively impacted the visual quality of lettuce. Liu et al. [26] highlighted that water-assisted UV treatments improved decontamination rates for blueberries contaminated with E. coli O157 compared to direct UV exposure. Martinez-Hernandez et al. [24] reported that Listeria monocytogenes was less sensitive to UVC, requiring 926 mJ/cm2 for effective disinfection. Chun et al. [28] observed that UVC doses between 100 and 800 mJ/cm2 reduced Listeria monocytogenes in fresh salads by 2.57 log CFU/g, showcasing significant inactivation at high doses. Kim et al. [30] explored Salmonella, Typhimurium, and Listeria inactivation on lettuce and found greater reductions when samples were irradiated from both sides and positioned closer to the UV source. Chang et al. [31] have determined the doses of UV light (254 nm) required for a 99.9% inactivation of the cultured vegetative bacteria, total coliforms, and standard plate count microorganisms and found that they were comparable. However, the viruses, the bacterial spores, and the amoebic cysts required about 3 to 4 times, 9 times, and 15 times, respectively, the dose required for E. coli. The order of microorganisms’ susceptibility to UV treatment is as follows: fungal spores < bacterial spores < mycobacteria < vegetative bacteria < viruses, according to Gryko et al. [32]. It is important to note here that microorganisms are unlikely to develop resistance to UVC radiation, according to published studies [14,33], despite some assertions to the contrary [34].
The standard germicidal effectiveness curve shown in Figure 1, which was first created by Gates [6] (red line) and later improved [35] (green line), is widely used to assess the potential of UV sources for germicidal applications. According to these curves, the maximum germicidal efficiency is for the UVC wavelengths in the range from 265 nm to 267.5 nm. If UVC radiation wavelengths are just around 30 nm shorter or longer than the wavelength with the highest efficiency, their effectiveness is reduced to half of their maximum value.

3. UVC Light Sources

The sun’s UVC light (<280 nm) is largely absorbed by molecular oxygen and the ozone layer before reaching the Earth [36]. For this reason, antimicrobial applications can only use artificial UVC sources, primarily gas discharge lamps, semiconductor light-emitting diodes (LEDs), and cathodoluminescent light sources (CL); for a review, see Ref. [37], for example.
UV light is often provided from xenon (190–1100 nm), deuterium (190–370  nm), and sealed mercury (253.7  nm) gas discharge lamps [38]. These sources need high operating voltages, are not practical for use because of their large size, and present adverse effects on the environment [38]. Additional drawbacks of current discharge lamps include their fragility, the need for an electronic ballast for operation, mercury toxicity, and the need for specialized materials such as quartz, all of which are currently being addressed by emerging technologies such as pulsed light and dielectric barrier discharge lamps [39]. Low-pressure mercury lamps (with a pressure of ~10 Torr), which account for about 90% of UV disinfection systems and have been used for over 90 years, are highly effective against microorganisms because their monochromatic emission is centered at 253.7 nm, which is very close to the peak of the germicidal efficiency curve (Figure 1). In recent years, these lamps have dramatically reduced mercury consumption, utilizing only 5 mg compared to prior models that used more than 30 mg [40]. Despite having various shortcomings, mercury lamps remain popular in the market owing to their availability. However, due to increased health and environmental concerns about mercury toxicity, the United Nations Minamata Convention [41] is implemented globally to fully ban or considerably decrease the use of mercury gas lamps. Two relatively important innovations are worth mentioning. High-power and high-voltage pulses are used with xenon-filled lamps to generate broad-spectrum white light from which UV can be extracted, while dielectric barrier discharge lamps produce a transient high-pressure glow [42].
The principle of operation of gas lamps containing excimer and exciplex molecules was known beginning in 1970, when the first excimer laser was developed [43]. Because they do not contain mercury, they have become increasingly popular as a replacement for low-pressure mercury lamps [44]. They produce high intensity quasimonochromatic radiation within the range of 100 nm to 350 nm as a result of the excited dimers’ (usually excited by electric discharge) spontaneous decay to their ground state. Different gas mixtures can produce radiation of different wavelengths. For example, the lamp exploiting excimer molecules generates radiation at the following wavelengths: Ar2*~146 nm, Kr2*~165 nm, and Xe2*~172 nm. The radiation wavelengths of lamps that work with exciplex molecules are ArBr*~165 nm, ArCl*~175 nm, KrI*~165 nm, ArF*~193 nm, KrBr*~207 nm, KrCl*~222 nm, KrF*~248 nm, XeI*~253 nm, XeF*~282 nm, and XeCl*~308 nm.
Semiconductors have been used as efficient light-emitting materials for electrically powered light-emitting diodes (LEDs), but the choice of semiconductors for UV light is limited by the necessity for wide band-gap structures (>3.1 eV) that have excellent crystal quality [45]. UVC semiconductor LEDs share the same technology concept as general lightning LEDs, and both are based on the III-nitride materials. However, in order to realize the UV emission, the bandgap of the semiconductor must be increased by substituting In for Al in InGaN (the semiconductor widely used for blue LEDs). The amount of Al required increases as the wavelength of the emission decreases. This presents a considerable technological challenge because Al is a smaller atom than In, and its incorporation causes significant lattice stress and produces structural defects due to its poor match with a crystal lattice. This, combined with challenges in p-type doping, results in relatively low UV LED efficiency [46]. As a result, the cost of producing semiconductor UVC LEDs is significantly higher than that of blue LEDs. On the other hand, UVC LEDs exhibit significant potential for antimicrobial applications due to their compact size and ability to emit at 265 nm (the emission wavelength achieved by optimizing the aluminum content in the semiconductor), where the GEC curve reaches its maximum value. Takano et al. [47] demonstrated an external quantum efficiency (EQE) of 20.3% for an AlGaN-based UVC LED emitting at 275 nm. The EQE reduces below 4% for LEDs emitting at wavelengths shorter than 265 nm owing to a substantial decrease in semiconductor crystal quality.
Cathodoluminescent UVC light sources, also known as electron-beam UVC light sources, have a simpler construction compared to UVC LEDs, with a basic thin-film anode and no p-type layer. An output power of 60 mW has been realized with electron-beam pulse-scanning pumping for a 270 nm emitting device constructed using AlGaN multi-quantum well heterostructures grown on c-Al2O3 substrates [48]. Kang et al. [49] demonstrated a CL device emitting at 246 nm (FWHM = 16 nm) with an output power of 430 mW via electron-beam pumping of a layer of YPO4: Bi3+ film. To compete with UVC LEDs, CL sources must be miniaturized in size [50], for example, by using field-emission cold cathodes. Watanabe et al. [51] fabricated a compact device (1.7 × 0.16 cm2 emitting surface) using a field-emission array as an excitation source and a hexagonal boron nitride as an emission phosphor. The device has a stable operation output power of 0.2 mW at 225 nm and a quantum efficiency of 0.6% with an excitation voltage of 8 kV. With an anode voltage of 9 kV and a current of 100 mA, the device achieved an output of 1 mW over several hours of operation.
Recent alternatives for creating UVC light sources include the use of non-linear optical (NLO) materials and LEDs, which utilize the blue-to-UVC upconversion from lanthanide-activated phosphors. A detailed discussion of the materials used for the later strategy is given below.

4. Short Primer on Lanthanide-Mediated Upconversion

Shorter-than-excitation wavelength (STEW) processes are light-emitting phenomena in which photons released by a material have greater energies (shorter wavelengths) than the original photons that produce them [52]. These processes are a contemporary approach to the production and utilization of light [53]. The primary STEW routes to light conversion are second-harmonic generation and anti-Stokes processes such as two-photon absorption, anti-Stokes Raman, and upconversion (UC). In UC processes, two or more low-energy photons are converted into one high-energy photon, typically through triplet-triplet annihilation and lanthanide-facilitated UC (LnUC), where the former has a better quantum efficiency and is mainly employed for visible-to-visible/UVA UC. LnUC is widely employed in infrared-to-visible UC and, to a lesser extent, visible-to-UVC/UVB/UVA UC [54].
Figure 2 illustrates the common LnUC processes and their efficiencies [55]. According to van der Ende et al. [56], the mechanism known as the APTE effect (Addition de Photon par Transferts d’Energie) [57,58] or ETU (energy transfer upconversion) is the most efficient of all. It may also include ground state absorption (GSA) succeeded by an energy transfer step, frequently denoted as GSA/APTE. The subsequent most effective mechanism is GSA, succeeded by excited state absorption (ESA), in a two-step absorption process. Cooperative sensitization (CS) and cooperative luminescence (CL) mechanisms involve one or more virtual energy levels and are less efficient than the previous two. Other relevant mechanisms are photon avalanche UC, energy migration-mediated UC, and donor–acceptor (D–A) energy transfer (a specific variant of APTE) [52,54,59].

5. Antimicrobial Applications of Lanthanide-Facilitated UC

Because visible light alone cannot effectively destroy microorganisms, near-infrared-to-visible and visible-to-visible UC are primarily used for antimicrobial applications in conjunction with photodynamic treatment (PDT). Antimicrobial photodynamic treatment (aPDT) works on the principle that a photosensitizer, often a photoactive dye, binds to the targeted cells and is triggered by light of a certain wavelength. The photosensitizer then creates reactive oxygen species (ROS), such as singlet oxygen (1O2), hydroxyl radical (OH), superoxide radical (O2•−), hydrogen peroxide (H2O2), etc., causing damage to the microorganism [60]. The first two are the most reactive and most cytotoxic species but have a short diffusion distance [61]. The mechanism underlying PDT antimicrobial action is illustrated in Figure 3 [62]. A photosensitizer is administered and attached to a microorganism, followed by exposure to light that provides suitable energy for its activation. Then, the photosensitizer transfers energy to surrounding oxygen to create ROS that destroy nearby microorganisms.
So far, aPDT has proven to be an effective treatment, in vitro and in vivo, for a wide range of microorganisms, including bacteria, fungi, viruses, and parasites, and is immune to resistance [61,63,64]. However, when aPDT is required to combat infections within tissues, it is impossible to activate photoactive dyes that are responsive to visible light, as tissue strongly absorbs external visible radiation. Considering that these dyes are among the most utilized photosensitizers, it is necessary to find a way to excite them at the location of their action. In contrast to red light, which can penetrate tissues up to 8–10 mm, and green and blue light to even shorter distances, near-infrared radiation at around 980 nm, commonly utilized to excite upconversion materials, achieves penetration depths of up to 10 cm in tissues [65,66]. Therefore, the visible light-responsive photosensitizers can be activated deep in tissues by attaching them to the near-infrared-to-visible UC nanoparticles. Then, near-infrared radiation is converted into visible light by UC nanoparticles, which subsequently activate photosensitizers. UC nanoparticles exploiting Yb3+ for near-infrared radiation absorption and Er3+ or Tm3+ for visible-light emission are typically used for this purpose, as shown in Table 1 [67]. This table also provides compositions of the employed UC nanoparticle-photosensitizer material for UC-facilitated aPDT, the UC host material, dopants involved in UC, absorption and emission wavelengths, and achieved germicidal effect. It is important to note that 980 nm excitation can lead to harmful heating effects because of significant absorption by water molecules, potentially resulting in severe damage to cells and biological tissues. The excitation at 808 nm, where Nd3+ absorbs based on its 4I9/24F5/2 transition, presents a promising alternative for near-infrared-to-visible upconversion due to the minimal absorption by water at this wavelength [68].

6. Lanthanide-Facilitated Near-Infrared-to-UV(C) and Visible-to-UV(C) UC

Near-infrared-to-UV and visible-to-UV LnUC can be realized with Ce3+, Er3+, Tm3+, and Tm3+/Gd3+, and Ce3+ activators via multiphoton processes (5-photon or more), typically through energy migration [78]/excitation energy-mediated cross-relaxation [79] or with Pr3+ through two-photon (visible-to-UV) or three-photon (NIR-to-UV) processes. The latter strategy, Pr3+-facilitated UC, is discussed in detail later in the article. The former strategies, which are referred to differently but follow similar principles, require host materials with core–shell or core–multishell morphologies, as well as careful control of dopants in different layers of the structure with high dopant concentrations in some circumstances. However, they are not optimal for antimicrobial applications as they primarily emit in the UVB/UVA spectral regions (except for the weak Gd3+ emissions in the 250–315 nm range) and are relatively inefficient due to the requirement of a large number of photons in the conversion process. The typical example of the energy migration UC process (Gd-mediated) is a UVB/UVA emission (300–370 nm) obtained by Yb3+ → Tm3+ → Gd3+ → Ce3+ energy transfer in the core–shell–shell nanostructure of NaYbF4:Gd/Tm (40/1%)@NaGdF4@CaF2:Ce (15%) [80].
Recently, Su et al. [81] demonstrated six-photon UC UV emission from Gd3+ (composed of emissions at 253 nm, 273 nm, 276 nm, 279 nm, 306 nm, and 311 nm) under 808 nm excitation using NaGdF4:49%Yb,1%Tm@NaGdF4:20%Yb@NaGdF4:10%Yb,50%Nd@NaGdF4 core–multishell nanoparticles. The mechanism is based on the energy transfer sequence Nd3+ → Yb3+ → Tm3+ + Yb3+ → Gd3+, as depicted in Figure 4. The 808 nm radiation excites Nd3+ sensitizers, which pass energy to a lower-energy Yb3+ excited state. This is followed by a sequence of five energy transfers from excited Yb3+ ions to the 3P2 state of Tm3+ with a subsequent non-radiative relaxation to the Tm3+ 1I6 level. The energy then passes to the Gd3+ 6P3/2,5/2,7/2 levels. Finally, energy transfer from Yb3+ involves the sixth photon to populate the Gd3+ 6DJ (J = 9/2, …, 1/2) levels from the 6P3/2,5/2,7/2 states.

7. Electronic and Optical Properties of Pr3+ Ions

Praseodymium (Pr) is one of the rare earth elements with atomic number 59. Its complete electron configuration in a neutral state is 1s22s22p63s23p63d104s24p64d105s25p64f36s2. In accordance with electron shell filling, the incomplete 4f electron shell—like in the case of other lanthanides—is shielded by the completely filled 5s25p6 shells, which considerably decreases the interaction of the 4f electrons with the nearest environment when Pr is incorporated into crystalline matrices. Its most stable oxidation state is +3, and the corresponding electron configuration is [Xe]4f2, where [Xe] stands for the electron configuration of xenon.
Various methods of the distribution of two electrons through seven f-orbitals (each orbital is doubly degenerated due to possible orientation of spin) lead to the formation of 91 microstates. Many of those 91 microstates have the same energy and are combined into seven LS terms: three spin-triplets, 3P, 3F, and 3H, as well as four spin-singlets, 1S, 1D, 1G, and 1I. According to Hund’s rule, the ground term is 3H. Spin–orbit coupling is important for the formation of the energy levels of heavy elements (located in the second half of the periodic table) since it is proportional to Zeff4 (Zeff < Z), where Z is the atomic number of the element. It splits the LS terms into J-manifolds, which numbers only 13 in the case of the Pr3+ ion: 3H4, 3H5, 3H6, 3F2, 3F3, 3F4, 3P0, 3P1,3P2, 1S0, 1D2, 1G4, and 1I6. The subscript denotes the values of the total momentum J, and in the crystal fields, each of the above-listed states can split into up to 2J+1 states, depending on the local symmetry of the crystal lattice site occupied by the Pr3+ ions.
Figure 5a shows the energy level scheme of a free Pr3+ ion 4f2 configuration calculated in Ref. [82] with the free ion Hamiltonian parameters from Ref. [83]. These free ion levels are split in the crystal fields, but the magnitude of such a splitting is not large and does not exceed a few hundred cm−1 for each J-manifold. This is why the free ion energy level schemes can be used for all lanthanide ions for the analysis and assignment of their spectra in crystals.
The first excited electron configuration of the Pr3+ ions is 4f5d, which corresponds to the excitation of one 4f electron into the outer 5d orbitals. This configuration is of the opposite parity to the 4f2 states: this is why the 4f–5d excitation (absorption) and 5d–4f (emission) transitions are parity allowed, have high intensity, and appear in the experimental spectra as broad bands. A total of 140 microstates of the 4f5d configuration give rise to 10 LS terms: five spin-triplets, 3P, 3D, 3F, 3G, and 3H, as well as five spin-singlets, 1P, 1D, 1F, 1G, and 1H. The inclusion of spin–orbit coupling in these states results in the formation of 20 J-manifolds: 3P0, 3P1, 3P2, 3D1, 3D2, 3D3, 3F2, 3F3, 3F4, 3G3, 3G4, 3G4, 3H4, 3H5, 3H6, 1P1, 1D2, 1F3, 1G4, and 1H5. The lowest excited level of the Pr3+ 4f5d configuration is 1G4 [84]. It is located at 61,171.9 cm−1 [84] and is strongly mixed with the 3H4 state from the same configuration located at 63,580.7 cm−1. Although 3H4 is indicated as the energetically lowest in the seminal work by Crosswhite et al. [85], the wave functions of these states calculated in Ref. [84] are 46% |1G4> + 49% |3H4> and 38% |1G4> + 50% |3H4> (only two greatest contributions to the wave functions are given, and all the remaining ones are omitted for the sake of brevity). Energy level assignment by the largest component of the wave function indicates that the |3H4> state produces the largest contribution to wave functions of both states, but since its contribution to the 63,580.7 cm−1 level is considerably greater than that of |1G4>, it is assigned here as the |3H4> state. Then, the ground state of the 4f5d configuration at 61,171.9 cm−1 is assigned to the |1G4> state; otherwise, two |3H4> states appear in the energy level scheme, and the |1G4> state simply vanishes. This is the standard approach to the energy levels assignment in the case of energy levels that are located closely in energy and whose wave functions are strongly mixed. Figure 5b shows the 4f5d levels of the Pr3+ ions [86]. The energy levels of the next excited electron configuration, 4f6s, start at around 100,000 cm−1.
When the lowest0energy 4f5d state is still higher in energy than the 4f2 1S0 level, Pr3+ ions exhibit several 4f2 → 4f2 emission transitions, which can be excited in different ways: either by resonant excitation to the emission levels or by non-radiative or down-conversion processes. One of those mechanisms can be UV excitation from the 3H4 ground state to the 1S0 level, which is followed by a downward cascade transition to the group of the closely located 3P0, 3P1,3P2, and 1I6 manifolds following emission transitions to the ground level 3H4 and nearest excited levels 3H5, 3H6, and 3F2.
A particular feature of the Pr3+ ions is that the lowest levels of the excited 4f5d electron configuration are located rather close to the 1S0 level ( 47,200 cm−1), which is the highest level of the 4f2 electron configuration. According to the NIST Atomic Spectra Database [86], the lowest state of the 4f5d electron configuration is at 61,171 cm−1 for the free Pr3+ ion, which indicates a strong mixture of states of both 4f2 and 4f5d configurations. The situation becomes even more complicated for Pr3+ ions in crystalline solids. Since the 5d states are spatially more extended, they interact strongly with the surroundings. In combination with a pronounced nephelauxetic effect (covalency), this can lead to a considerable red shift (decrease in energy) of the 4f5d states (Figure 6a), often even below the 1S0 level. Such a situation is favorable for obtaining broad UV emission from the Pr3+ 4f5d→4f2 (Figure 6b) if the Stokes shift of this emission is smaller than 3000 cm−1 [87,88].
According to the crystal field theory, the energy of d-orbital splitting depends on the anion charge/anion radius (spectrochemical series): I < Br < Cl < S2− < F < O2− < N3− < C4−, symmetry (coordination number and site symmetry): octahedral > cubic, dodecahedral, square antiprismatic > tetrahedral, and Pr3+-ligand distance: Dq = 35 Z e / 4 R 5 (R is the cation–anion distance, Z represents the valency of anions, and e is the electron charge) [89]. The covalency of the Pr3+–ligand bonds mainly depends on the polarizability (type) of the anions involved (sulfides > nitrides > oxides > fluorides) and charge density on the surrounding anions (aluminates and gallates > silicates and germanates > borates > phosphates > sulfates) [89]. The increase in the covalency of the Pr3+–ligand bonds leads to a decrease in 4f5d energy.
Upon excitation of Pr3+ into the 4f5d configuration, a series of emission transitions from that state to the 3HJ (J = 4, 5, 6) and 3FJ (J = 2, 3, 4) levels occurs. These transitions are parity allowed, so they provide faster, broader, and significantly more intense emissions (over the 200–400 nm spectral range, depending on the host) compared to the narrow-line emission from parity-forbidden 4f2 → 4f2 transitions (in the VIS and NIR spectral range). If the lowest energetic 4f5d state and 1S0 level are close in energy, emissions from both 4f5d → 4f2 and 4f2 → 4f2 transitions occur. Table 2 provides the wavelengths of emissions from the Pr3+-based 4f5d spin-allowed transitions in several host materials along with the description of their structure.
Much more comprehensive data can be found in Refs. [90,91]. In instances lacking literature data for the energy of the Pr3+ first spin-allowed transition, it can be estimated using an extensive dataset available for Ce3+ or from the relatively limited literature data for Nd3+, using the following equation [90,91]:
E P r 3 + E C e 3 + + 12,240   c m 1 E N d 3 + 10,460   c m 1 .

8. Mechanisms of Visible-to-UV Upconversion in Pr3+

Historically, the first UC processes in Pr3+ have been demonstrated through observation of blue (3P03H4) and red (1D23H4) emissions under near-infrared excitation (NIR-to-VIS UC) and blue emission under red excitation (VIS-to-VIS UC) [128,129,130,131,132,133,134,135]. The quantum yields of these processes were less than 1% [136,137].
Upconversion emission from the Pr3+ 4f5d state occurs under blue excitation via GSA/ESA (Figure 7a) or GSA/APTE (Figure 7b) processes. Except for low Pr3+ concentrations, the former process is more efficient than the latter [138]. Both kinds of processes can take place at the same time, and the formula to determine the ratio of APTE to ESA rates in the UC process was given by Sun et al. [138]. The presence of a delay in the transient of the emitted signal (i.e., the rise in emission before the decay) shows the presence of an energy transfer (APTE), allowing time-resolved emission measurements to be utilized to determine the dominant UC process. Both GSA/ESA and GSA/APTE are two-photon processes and use either 3P0 (blue arrows) or 1D2 (red arrows) levels as an intermediate level. For the efficient GSA/ESA UC, a long-lived intermediate state is generally required. Consequently, effective GSA/ESA UC via the 3P0 level requires hosts with low phonon energies for minimizing non-radiative transitions to the 1D2 level (3P01D2 multiphonon relaxation and [3P0,3H4] → [1D2,3H6] cross-relaxation). The 1D2 level exhibits a significantly longer radiative decay time than 3P0 since a multiphonon relaxation to the adjacent lower energetic 1G4 is weak because of a large energy gap of around 6500 cm−1 and a weak radiative transition to the 3H4 ground level (this transition is spin-forbidden). In 1D2 level-mediated UC, obtaining a strong population of the 1D2 level through the 3P0 level is one of the essential conditions to achieve efficient UC, as direct excitation from the 3H4 to the 1D2 level is weak. This can be realized in high-energy phonon hosts that promote strong multiphonon relaxation from 3P0 to 1D2 or in hosts with a metal-to-metal charge transfer band with energy adequate to bridge 3P0 to 1D2 states.
For a free Pr3+ ion, the lowest energy 4f5d state is 1G4 [86]. In Pr3+-activated LaCl3, as previously stated, the wave function of 4f5d 1G4 state is strongly mixed by a spin–orbit interaction in fractions of 46% 1G4 (spin singlet) + 49% 3H4 (spin-triplet) [84]. The wavefunction of the 4f5d 3H4 state is also spin–orbit-mixed 50% 3H4 + 38% 1G4 [84]; therefore, in both cases, the spin-triplet state dominates, as expected according to Hund’s rules. The degree of mixing between 1G4 and 3H4 states may differ across various hosts; however, the fast emission decays and broad emission bands that originate from the lowest energetic 4f5d state to 3H4,5,6 spin-triplet states indicate the spin-allowed nature of these transitions and the predominant role of the 3H4 spin-triplet state in the wavefunction of the lowest energy 4f5d state. Therefore, one may argue that both 3P0 (4f2) → 4f5d and 1D2 (4f2) → 4f5d transitions are only partially spin-allowed and that the efficiency of 1D2-mediated UC depends on the amount of singlet state contribution to the wavefunction of the lowest energy 4f5d state. In this sense, 4f5d → 3H4,5,6 may also be characterized as partially spin-allowed. It should be noted that for some hosts, UC via the 1D2 state is not possible if the energy gap between the 1D2 level and the 4f5d states is greater than the excitation energy to the 3PJ states.
The theoretical maximum of UC quantum yield is 50% [139] due to the absorption of two photons per one emitted. However, most of the Pr3+ UC materials show significantly smaller UC quantum yields of around 0.01%, which are also substantially lower than the yields of NIR-to-visible lanthanide UCs (quantum yields up to 10% can be achieved in β-NaYF4:2% Er3+,18% Yb3+ [140,141]). For example, previous work [142] indicates a UC quantum yield of 0.0019% (excitation power flux of 1.65 mW cm−2) for X2-Y2SiO5:Pr3+,Li+, while 3.9-fold stronger UC emission has been observed in β-Y2Si2O7:Pr3+ [143]. Recent studies on bromide-based [144] and borate-based materials [145] have demonstrated significantly improved UVC UC. Furthermore, [145] is a good example of a strategy using high-phonon-energy materials to populate the 1D2 level. UC efficiency is mainly affected by a relatively small absorption cross-section of the 3P0,1,2 levels [146] and emission quenching due to non-radiative multiphonon relaxations, cross-relaxations, energy transfer/migration processes (such as Pr3+–Pr3+ interactions), and Pr3+ ion interactions with species attached to phosphor particle surfaces [147]. The process of multiphonon relaxation from the 3P0 level to the 1D2 level is a key reason why the UC efficiency using 3P0 as an intermediate level is limited. The energy difference between these two levels ( E ) is around 3800 cm−1, so it can be easily bridged by several high-energy phonons. The multiphonon relaxation rate (WMPR) for an excited state resulting from stimulated phonon emission can be derived from the energy-gap law [148]:
W M P R   = β · exp α · ( E 2 ω m a x )   ,
where ω m a x is the cutoff phonon energy (maximum optical phonon energy) of the host, and α and β are the characteristic coefficients. For many silicates, which usually have a cutoff phonon energy of about 1080 cm−1, it takes only three to four phonons to depopulate the 3P0 level, with a multiphonon relaxation rate of about 28,500 s−1 (the non-radiative lifetime of around 35 μs; calculated using α = 4.7 × 10 3 s−1 and β = 9 × 10 7 cm−1 [147]). For this reason, high-energy phonon hosts, such as silicates and phosphates (see Table 2), are not well-suited for the UC exploiting 3P0 as an intermediate level. On the other hand, using high-phonon-energy materials to populate the 1D2 level has recently been demonstrated as an perspective strategy to increase the efficiency of UC using the 1D2 intermediate level [145]. At high Pr doping concentrations, usually ≥0.5 mol%, the energy transfer process becomes efficient because of the reduced Pr3+–Pr3+ distance, and the UC process is hindered by cross-relaxations due to enhanced non-radiative coupling between ions [145,149]. The situation with nanoparticles is somewhat more complicated. Due to phonon confinement, the cutoff phonon energy may be smaller than the maximum optical phonon energy in the bulk. On the other hand, Pr3+ ions interact more strongly with species at particle surfaces, for example, OH groups, resulting in the emission quenching due to this interaction. These complex nature of excitation and emission transitions explains the low efficiency of Pr3+ two-photon blue-to-UV UC compared to efficiencies of lanthanide-facilitated near-infrared-to-visible UC. However, it is worth noting that substantial research into efficient lanthanide near-infrared-to-visible UC materials has been going on for several decades. Over time, many strategies for increasing UC efficiency have been developed, such as activator concentration optimization; improved material preparation methods, including materials with core/shell morphologies; passivation of particle surfaces; the use of plasmon particles and dyes to enhance absorption, etc. Such efforts have only recently started with Pr3+ blue-to-UV UC materials. The Li-codoping of Y2SiO5:Pr3+ resulted in enhanced upconversion emission relative to the non-codoped material, attributed to an increased particle size (flux effect) [126,142]. As the size of crystals enlarge, the number of Pr3+ ions in the vicinity of the surface of the particles reduces, thus limiting non-radiative losses [11]. Malyukin et al. [150] suggested that Li+-codoping of Y2SiO5:Pr3+ inhibits the clustering of Pr3+ ions within the host, hence reducing unwanted cross-relaxation depopulation of the 1D2 level. However, the mechanism responsible for the improvement in UC efficiency remains ambiguous according to Cates et al. [126]. Apparent strategies for the improvement in Pr3+ blue-to-UV UC can be the coupling of the excitation process with the localized surface plasmon resonance (LSPR) of noble metals or the implementation of dye sensitization to enhance the absorption efficiency of UC materials, as molecular dyes possess significantly greater absorption cross-sections (10−17–10−16 cm2) in contrast to the 4fn-4fn absorptions of lanthanides (<10−20 cm2) [151], similarly to the strategies employed with the NIR-to-VIS UC materials [59]. However, these additive materials show high UV absorption, implying that the resultant UC will be absorbed by them. UC efficiency may be improved by coupling the emission with LSPR of UV plasmonic metals, such as Al, Ga, and Rh, but this is yet to be tested.

9. Antimicrobial Applications of Lanthanide-Facilitated Visible-to-UVC UC

Table 3 shows a selection of materials doped with Pr3+ that have been used to generate UVC UC emission that resulted in a biocidal effect. Zhao et al. [152] conducted in vivo animal studies in addition to the standard practice of demonstrating the germicidal effects of Pr3+ UVC emission by exhibiting the findings of in vitro antibacterial research. The authors developed the antibacterial wound dressing in the form of a polymeric antibacterial composite film that is composed of polyvinyl alcohol, sodium alginate, and Y2SiO5: Pr3+. The film was successfully used to inhibit bacteria in actual wounds, as illustrated in Figure 8.
Falat et al. [157] conducted Tm3+ co-activation of Pr3+-activated Y2Si2O7 powders to achieve a 370 nm UVA up-conversion emission (Tm3+), alongside Pr3+ emissions in the UVC (278 nm) and UVB (308 nm) spectrum regions. The authors evaluated the biocidal efficacy of materials subjected to 447 nm laser irradiation on biofilms formed by A. baumannii, S. aureus, and C. albicans (Figure 9a). They discovered that the application of Y2Si2O7:Pr3+,Tm3+,Yb3+ powder as a UV light source resulted in a reduction in biofilm-forming microorganism viability to 45.5 ± 2.5% (S. aureus), 39.0 ± 3.0% (A. baumannii), and 36.5 ± 2.5% (C. albicans), significantly surpassing the results obtained with sole Pr3+ doping (approximately 25–35%), as illustrated in Figure 9b). The authors observed an elevation in the concentration of reactive oxygen species (ROS) following 10 min of irradiation of the biofilm, in contrast to the control sample (biofilm without irradiation), as depicted in Figure 9c). The observed increase in the ROS concentration (~54 ± 2%) in Y2Si2O7:Pr3+ powders led the research group to conclude that the phototoxicity mechanism of only Pr3+-doped materials include both DNA damage and ROS production.
Recent reports indicate that Pr3+ UVC emissions obtained through upconversion can be effectively employed for a range of important applications in addition to antimicrobial use. For example, Li2SrGeO4:Pr3+ is used for UVC optical marking, including static and dynamic labeling [159], and the upconversion emission of YOBr:Pr3+/polydimethylsiloxane is employed to induce photocatalytic water splitting via NiO-loaded NaTaO3:La [144].
The emission from Pr3+ does not necessarily need to be produced through upconversion for it to be applicable in sterilization applications. Zhang et al. [160] demonstrated 24 h of continuous 222 nm UVC persistent luminescence in X-ray excited Sr2P2O7:Pr3+ phosphor. This phosphor was able to effectively inactivate infectious methicillin-resistant Staphylococcus aureus (MRSA) within 30 min in an excitation-free manner. Cheng et al. [161] demonstrated that the inactivation of Staphylococcus aureus bacteria under 254 nm mercury lamp irradiation is substantially enhanced when the bacteria are irradiated through PDMS film containing 265 nm-emitting Ba2MgSi2O7:0.4%Pr3+ phosphor particles.
It is important to point to the spectral overlap between 4f5d emission spectra with the germicidal efficiency curve (GEC), Figure 10, as the criteria for assessing the potential of the Pr3+ UC material for antimicrobial use in addition to its quantum efficiency. Figure 10 shows that only a fraction of the LaPO4:Pr3+ 4f5d-based emission (28.2%) falls into the GEC spectral region (green-shaded), whereas much of the Lu7O6F9:Pr3+ 4f5d-based emission (67.4%) falls in the GEC region with the highest efficiency.
Here, we propose a spectral overlap coefficient (SOC) as one of the key indicators of the Pr3+ UVC-emitting materials’ germicidal potential in the following form:
SOC = λ m i n λ m a x G E C λ · I n λ ·   d λ λ m i n λ m a x G E C λ · d λ ,
where λ represents the wavelength, I n λ is the normalized emission intensity of the material (normalized to the maximum intensity value of 1), and G E C λ stands for the germicidal efficiency curve. Table 4 provides the SOC values for Pr3+ materials that are displayed in Figure 10. The value of germicidal potential (GP) of the material can be used for comparison between different Pr3+ activated materials:
GP = SOC × EQE ,     or alternatively           GP = SOC × Emission power flux Excitation power flux ,
where EQE is the external quantum efficiency of the material. It is important, however, to remember that GEC is mainly derived from the UV light germicidal effect on E. colli bacteria, so it can take different values and spectral shapes for different microorganisms, as discussed in Section 2. Furthermore, the GEC does not cover wavelengths below 240 nm, where a further decrease its value is not necessarily guaranteed to happen. While deriving his germicidal effectiveness curve, Gates [6] found that the germicidal effect on E. colli starts to increase with a decrease in the UV light wavelength below 240 nm. It is therefore important to derive the GEC curve for wavelength values less than 240 nm in the future. Additionally, we wish to highlight that GEC can vary depending on the specific microorganism, as noted in Section 2. Therefore, the SOC needs to be adjusted for other pathogens, such as C. albicans or MRSA, by deriving and utilizing microorganism-specific GECs.

10. The Toxicity of Some Elements Frequently Used for Lanthanide Upconversion Materials

The toxicity of elements commonly employed in lanthanide UC materials raises serious concerns about their environmental and health effects [162,163]. Like other chemicals or pharmaceuticals, properties that are advantageous from optical and biomedical standpoints can also lead to unexpected potentially hazardous toxicities [164]. The issue is particularly significant with nanoparticles [165], which exhibit greater toxicity to human health compared to larger particles of the same chemical substance; it is commonly proposed that toxicity levels are inversely related to the size of the nanoparticles [166,167,168]. As researchers look for safer alternatives, thorough investigations of these materials’ biocompatibility and ecological footprint become increasingly important. Moreover, it is critical to raise awareness of the potential risks associated with the preparation and use of these products.

10.1. Yttrium

Occupational exposure to yttrium has been documented in workers involved in electronic waste recycling. These individuals exhibit increased levels of yttrium in their blood and urine compared to the general population. In one study, Y levels were reported at 10–15 μg/L in urine, significantly higher than those observed in the general population (5 μg/L) [169]. This suggests the accumulation of this element in the body due to repeated exposure. Yttrium oxide (Y2O3) has been shown to induce apoptosis and necrosis in HEK293 cells (human embryonic kidney cells) through mechanisms involving elevated levels of reactive oxygen species (ROS) and disruption of mitochondrial function [170]. This element affects the integrity of the mitochondrial membrane and causes oxidative damage to DNA. In human endothelial cells exposed to Y2O3 nanoparticles (concentrations of 10 μg/mL), a 40% increase in IL-8 and ICAM-1 levels was observed compared to the control group [171]. This may contribute to the development of chronic diseases, especially in cases of long-term exposure [171]. Liu et al. [172] mention an association between yttrium exposure and hormonal level changes, such as a 20% decrease in TSH levels in infants exposed during pregnancy, indicating endocrine system disruption. In a study on smokers, yttrium concentrations in sperm samples were reported at 1.5–2 μg/L, significantly higher than those detected in non-smokers (approximately 0.8 μg/L). Prolonged exposure was correlated with increased DNA fragmentation in sperm and reduced sperm motility [172]. Prolonged exposure to yttrium in rats causes testicular damage, reducing sperm quality and serum testosterone while increasing cellular apoptosis and cytosolic Ca2+ levels [173].
In Y2SiO5-type compounds, it is likely that, in biological systems, due to the complexation of Y3+ with amino acids containing COO groups and the exposure of particles in the acidic environment of the stomach, their surface becomes enriched in silicon. This may reduce further solubilization of the particles, decreasing the toxicity of Y3+ and the rare earth elements with which the compound is doped. The same does not apply to fluorides, which can solubilize to a much greater extent in the human digestive system.

10.2. Gadolinium

The toxicity of gadolinium is well documented, as it is administered as a contrast agent for MRI. It has been observed that Gd can remain in bone and brain tissues for up to 8 years after exposure [174]. Gadolinium-based contrast agents (GBCA) can induce acute kidney injury, especially in patients with chronic kidney disease. In patients with diabetic nephropathy, the risk of acute kidney injury increased by 60% after the use of GBCA [175]. Nephrogenic systemic fibrosis (NSF) has been reported as a severe adverse reaction in patients with stage 5 chronic kidney disease. Studies have indicated that gadolinium administered in contrast agents, such as gadodiamide, can trigger this condition [176,177]. Gadolinium oxide nanoparticles (Gd2O3) induced cytotoxicity in human umbilical vein endothelial cells (HUVEC). At a concentration of 50 μg/mL, Gd2O3 caused lipid peroxidation and increased ROS, mitochondrial dysfunction, and apoptosis. This underscores the high toxicity of Gd at elevated concentrations [178]. GBCAs have been reported to induce chronic pain, including fibromyalgia, after repeated administrations of Gadovist, a GBCA agent [179]. In other studies, Omniscan (another GBCA) increased the levels of matrix metalloproteinase-1 (MMP-1) and the tissue inhibitor of metalloproteinase-1 (TIMP-1) in human dermal fibroblasts [180]. In vitro studies have shown that gadolinium can stimulate fibroblast proliferation and increase hyaluronic acid production without affecting collagen synthesis. This suggests that Gd may contribute to fibro-trophic processes in tissues [181].

10.3. Erbium

In industries using erbium (e.g., manufacturing optical components), workers may be exposed to elevated levels of dust containing erbium. In one study, erbium levels in industrial dust were estimated at 3–5 mg/m3, and prolonged exposure was associated with respiratory irritation and chronic inflammation [182]. Erbium-doped nanoparticles induced apoptosis and necrosis in human bronchial epithelial cells (BEAS-2B). Studies showed that YAl3(BO3)4 (YAB) nanoparticles doped with erbium increased ROS levels and affected mitochondrial integrity at concentrations of 20–50 μg/mL [182,183]. In vitro studies indicated a decrease in cell viability by over 30% following prolonged exposure (48–72 h) to YAB nanoparticles doped with erbium. These effects were more pronounced in pulmonary cells compared to other cell lines. Erbium-based materials used in dentistry have not shown significant cytotoxic effects on healthy tissues in routine clinical applications. However, studies recommend avoiding repeated exposure in industrial environments [184].

10.4. Lutetium

Lutetium-177 (177Lu) is used in radiotherapeutic treatments for cancer, particularly in peptide receptor radionuclide therapy (PRRT). In these applications, toxicity is primarily associated with the emitted radiation. The most common adverse effects include myelosuppression (a reduction in blood cell counts) and renal toxicity, observed in patients exposed to doses of 5.55 GBq (177Lu) per treatment [185]. Lutetium is reported to be relatively chemically inert in its non-radioactive forms; however, studies have shown that lutetium-doped nanoparticles can provoke oxidative effects in exposed cells. For instance, lutetium-doped borate nanoparticles induced a decrease in the viability of human lung cells (A549) by approximately 20% at a concentration of 50 μg/mL through oxidative mechanisms [183]. In preclinical studies, lutetium-doped nanoparticles used in radiotherapeutic therapies demonstrated a low toxicity profile in normal cells. However, their chronic use requires further studies to assess long-term risks [184]. Workers in industries handling lutetium, such as laser manufacturing or materials for radiotherapy, may be exposed to aerosols or dust containing this element. Compared to La, the lightest lanthanide, Lu is 200 times more toxic [186]; studies have reported lutetium concentrations of up to 2–3 mg/m3 in these environments, but the direct toxic effects on health have not been well-documented [187].

10.5. Thulium

Thulium (Tm) is one of the least studied rare earth elements. The available literature identifies health effects primarily related to its nanotechnological and medical applications. Tm3+-activated nanoparticles have been used in radiotherapy to enhance the efficiency of treatments against metastases. In industrial settings, repeated exposure to Tm can occur through the handling of materials containing this element. Reported levels in industrial air range between 1 and 3 mg/m3. Direct health effects on workers have not been detailed, but it is considered that Tm, similar to other rare earth elements, may contribute to respiratory irritation and chronic inflammation with prolonged exposure [188]. Adverse effects include oxidative stress and changes in gene expression associated with apoptosis, but only at high doses. In therapeutic applications, Tm-activated nanoparticles have proven effective and relatively safe for healthy tissues [189]. Studies have shown that thulium nanoparticles can induce apoptosis in cancer cells, such as cutaneous squamous cell carcinoma, at concentrations of 50 μg/mL. In normal cells, cytotoxicity was minimal, suggesting a promising potential for selective therapeutic treatments [189]. At high exposures, Tm3+-activated nanoparticles can cause oxidative stress by increasing ROS levels. In experiments on bronchial epithelial cells (BEAS-2B), Tm3+-activated nanoparticles reduced cell viability by 15–20% within 24 h, suggesting oxidative effects at prolonged exposures [190].

10.6. Praseodymium

Praseodymium (Pr) is used in various technological applications, including alloys, lasers, and optical materials. Studies on its toxicity have highlighted oxidative effects and its cytotoxic potential, particularly in cases of occupational exposure and contaminated environments. Praseodymium, similar to other rare metals, exhibits low to moderate toxicity. The ingestion of soluble praseodymium salts poses mild toxicity, whereas insoluble salts are non-toxic. These substances are irritants to the skin and eyes. Praseodymium poses significant hazards in occupational settings, primarily because its dust and gases can be inhaled.
Praseodymium nanoparticles (Pr2O3) have demonstrated significant cytotoxicity in human lung cells (A549). Research has shown a reduction in cell viability by approximately 25% at a concentration of 50 μg/mL due to increased ROS levels and mitochondrial dysfunction [185]. In vitro experiments revealed that Pr caused oxidative DNA damage in bronchial epithelial cells, contributing to apoptosis through the generation of reactive oxygen species. These effects were more pronounced with prolonged exposure, exceeding 24 h, to Pr2O3 nanoparticles [191,192].
Workers in industries involved in electronic waste recycling or alloy manufacturing containing praseodymium are exposed to dust or aerosols containing this element. Reported levels in industrial air range from 2 to 4 mg/m3, and chronic exposure has been associated with respiratory symptoms such as chronic irritation and inflammation [193] Experimental studies suggest that the accumulation of praseodymium in the body can affect the central nervous system by disrupting normal neuronal functions, although the exact mechanisms remain unclear. In animal models, accumulation in the cerebral cortex has been linked to increased oxidative stress [194].
Pr is also used in pesticides and fertilizers, leading to soil and water contamination. Exposure through the food chain can result in accumulation in the human body, with some studies suggesting it may contribute to reduced total protein and serum albumin levels in exposed populations [195].
In biological experiments utilizing UPM doped with Pr3+, it is crucial to consider the toxicity of the elements on cell lines to prevent the potential synergistic effects of metal cation toxicity and ultraviolet light emitted by the material. Pr exhibits notable antibacterial activity by disrupting bacterial membrane integrity and permeability, thereby inhibiting bacterial survival rates. The primary mechanism involves ion exchange and the production of reactive oxygen species (ROS), which collectively impair normal cellular functions. The interaction of praseodymium ions (Pr3+) with bacterial membranes results in structural destabilization and increased permeability. Studies on Escherichia coli have demonstrated that Pr3+ displaces calcium ions (Ca2+) from their membrane binding sites due to similar ionic radii, thereby exacerbating membrane instability and promoting ion leakage [196,197]. Pr3+ ions induce ROS generation within bacterial cells, leading to oxidative stress and eventual cell death. Research on aquatic microorganisms, including Vibrio fischeri and Tetrahymena thermophila, highlights Pr3+ toxicity across a range of effective concentrations (3.5–21 mg/L), with cell death being directly linked to oxidative damage [198].

10.7. The Toxicity of Fluorides

Fluoride is a chemical ion naturally present in water, soil, food, and dental hygiene products, such as toothpaste. While fluoride has well-known benefits in preventing dental cavities, excessive exposure can have toxic effects on human health. The acute toxicity of fluoride is due to its ability to form complexes with Ca ions in the body, leading to severe hypocalcemia, which can impair the normal functioning of the nervous and muscular systems [199]. Acute exposure to high concentrations of fluoride, either through accidental or intentional ingestion, can cause nausea, vomiting, abdominal pain, diarrhea, and, in severe cases, respiratory or cardiac failure. The median lethal dose (LD50) for sodium fluoride is approximately 5 mg/kg body weight for an adult [199]. Chronic exposure to excess fluoride (concentrations above 1.5 ppm in drinking water) can lead to dental fluorosis [200]. Prolonged consumption of water with high fluoride levels (over 4 ppm) can cause skeletal fluorosis, characterized by the accumulation of fluoride in bones, making them denser but more brittle [201]. Some epidemiological studies suggest that chronic fluoride exposure may impact cognitive development. A study in China found that children from areas with high fluoride levels in water had lower IQ scores compared to those from areas with normal levels [202]. Fluoride interferes with thyroid function by inhibiting the conversion of T4 to T3, which can lead to subclinical hypothyroidism. This effect has been observed particularly in populations exposed to high fluoride levels [203]. Long-term exposure to high doses of fluoride can affect kidney function, especially in individuals with chronic kidney disease. Fluoride can reduce the ability of kidneys to excrete this ion, leading to its accumulation in the body [204]. Other alkali metal ions, such as Cs+, Li+, and Na+, do not exhibit significant toxic effects at the low concentrations that may arise in industrial applications using UC materials. However, the presence of these elements in chemical compounds is associated with high solubility, necessitating the integration of such materials into polymer composites, microencapsulation, or coatings of particles with other low-solubility inorganic materials.

11. Conclusions

Because of its unique features, UVC radiation has found important and contemporary applications, for example, in the decontamination and disinfection of environments, water, and food; in cancer therapy; in photocatalysis; and for invisible identification tags. Some Pr3+-activated materials can generate intense UVC light from their excited 4f5d states through mechanisms of scintillation, downshifting, cathodoluminescence, and upconversion, provided that the lowest energetic 4f5d state has a lower energy than the 1S0 level of the 4f2 electron configuration. When excited by blue light into the 3PJ levels, Pr3+ can transfer the energy of two blue photons into one UVC photon via the GSA+ESA or GSA+ETU process, using 3P0 or 1D2 level as an intermediate state. It is not uncommon for all processes to take place concurrently. Blue-to-UVC UC opens up the opportunity for utilizing easily available and inexpensive blue light sources for the aforementioned applications. Although rather inefficient, with values of quantum yields well below 1%, Pr3+ blue-to-UVC UC materials have already demonstrated exceptional antimicrobial efficiency, as reported in a number of publications. Their biocidal efficacy can be enhanced by coupling them with inorganic photocatalysts, such as ZnO, TiO2, and BiOCl, for antimicrobial photodynamic treatment, analogous to the application of near-infrared-to-visible upconverters with visible-excitable photosensitizers. The alternative approach for UVC generation by lanthanide upconversion materials is mainly via Yb3+ to Tm3+ energy transfer-mediated upconversion, which is highly inefficient since it requires five near-infrared photons to generate one UVC photon. The low upconversion efficiency of Pr3+ upconversion materials is a consequence, however, of much less work devoted to the development of these materials compared to work committed to the development of near-infrared-to-visible upconverters. The UVC generation efficiency can be improved by (1) careful selection of host materials for Pr3+ that considers the matching of excitation energy with energies of 3H43PJ absorptions and appropriate phonon spectrum, (2) improving synthesis methods to obtain well-crystalline material, and (3) using charge compensation where appropriate. To enhance the germicidal efficiency of the material, it is crucial to achieve a substantial overlap between the 4f5d emission spectrum and the germicidal efficiency curve alongside the enhancement of upconversion quantum efficiency. This can also be accomplished by selecting appropriate hosts for Pr3+. Finally, it is essential to raise awareness regarding the toxicity of materials used in the production of upconversion materials, especially nanoparticles, when real-case applications are considered.

Author Contributions

Conceptualization, M.D.D.; methodology, M.D.D., R.B., M.G.B. and M.S.; formal analysis, M.D.D., Ž.A., T.F. and M.G.B.; investigation, C.M., R.B., T.D., Ž.A., T.F., M.D.D. and G.D.D.; data curation, C.M., R.B., T.D., Ž.A., T.F., Z.R. and G.D.D.; visualization, Ž.A., Z.R. and M.D.D.; writing—original draft preparation, M.D.D., R.B., M.G.B. and T.D.; writing—review and editing, M.D.D., Ž.A. and M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Romania’s National Recovery and Resilience Plan, PNNR [project grant number C9-I8-28/FC 760107/2023].

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

This study was supported by Romania’s National Recovery and Resilience Plan, NRRP [project grant number C9-I8-28/FC 760107/2023]. The authors from the Vinča Institute of Nuclear Sciences would like to acknowledge funding from the Ministry of Science, Technological Development, and Innovation of the Republic of Serbia under contract 451-03-136/2025-03/200017. The authors from Heinrich Heine University Düsseldorf gratefully acknowledge a materials cost allowance from the Fonds der Chemischen Industrie e.V; a scholarship from the “Junges Kolleg” of the North-Rhine Westphalian, Academy of Sciences and Arts; and generous funding by the Boehringer Ingelheim Foundation. The authors from Heinrich Heine University Düsseldorf and Vinča Institute of Nuclear Sciences are thankful for a Project-Related Personal Exchange of the German Academic Exchange Service (grant No. 57657161).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kowalski, W. Ultraviolet Germicidal Irradiation Handbook: UVGI for Air and Surface Disinfection, 1st ed.; Springer: Berlin/Heidelberg, Germany, 2009; ISBN 978-3-642-01998-2. [Google Scholar]
  2. Moore, G.; Ali, S.; Cloutman-Green, E.A.; Bradley, C.R.; Wilkinson, M.A.; Hartley, J.C.; Fraise, A.P.; Wilson, A.P.R. Use of UV-C Radiation to Disinfect Non-Critical Patient Care Items: A Laboratory Assessment of the Nanoclave Cabinet. BMC Infect. Dis. 2012, 12, 174. [Google Scholar] [CrossRef]
  3. Cieplik, F.; Deng, D.; Crielaard, W.; Buchalla, W.; Hellwig, E.; Al-Ahmad, A.; Maisch, T. Antimicrobial Photodynamic Therapy—What We Know and What We Don’t. Crit. Rev. Microbiol. 2018, 44, 571–589. [Google Scholar] [CrossRef]
  4. Rezaie, A.; Leite, G.G.S.; Melmed, G.Y.; Mathur, R.; Villanueva-Millan, M.J.; Parodi, G.; Sin, J.; Germano, J.F.; Morales, W.; Weitsman, S.; et al. Ultraviolet A Light Effectively Reduces Bacteria and Viruses Including Coronavirus. PLoS ONE 2020, 15, e0236199. [Google Scholar] [CrossRef]
  5. Arrage, A.A.; Phelps, T.J.; Benoit, R.E.; White, D.C. Survival of Subsurface Microorganisms Exposed to UV Radiation and Hydrogen Peroxide. Appl. Environ. Microbiol. 1993, 59, 3545–3550. [Google Scholar] [CrossRef] [PubMed]
  6. Gates, F.L. A Study of the Bactericidal Action of Ultra Violet Light. J. Gen. Physiol. 1930, 14, 31–42. [Google Scholar] [CrossRef] [PubMed]
  7. Rastogi, R.P.; Richa; Kumar, A.; Tyagi, M.B.; Sinha, R.P. Molecular Mechanisms of Ultraviolet Radiation-Induced DNA Damage and Repair. J. Nucleic Acids 2010, 2010, 592980. [Google Scholar] [CrossRef]
  8. Fondriest Environmental, Inc. Solar Radiation and Photosynethically Active Radiation. Fundamentals of Environmental Measurements. Available online: https://www.fondriest.com/environmental-measurements/parameters/weather/photosynthetically-active-radiation/ (accessed on 3 March 2025).
  9. Du, Y.; Ai, X.; Li, Z.; Sun, T.; Huang, Y.; Zeng, X.; Chen, X.; Rao, F.; Wang, F. Visible-to-Ultraviolet Light Conversion: Materials and Applications. Adv. Photonics Res. 2021, 2, 2000213. [Google Scholar] [CrossRef]
  10. Wang, X.; Mao, Y.; Wang, X.; Mao, Y. Recent Advances in Pr3+-Activated Persistent Phosphors. J. Mater. Chem. C Mater. 2022, 10, 3626–3646. [Google Scholar] [CrossRef]
  11. Li, L.; Zi, L.; Yang, F.; Feng, S.; Wang, C.; Yang, Y. Pr3+-Based Visible-to-Ultraviolet Upconversion. A Minireview. Adv. Phys. Res. 2025, 4, 2400097. [Google Scholar] [CrossRef]
  12. Besaratinia, A.; Yoon, J.; Schroeder, C.; Bradforth, S.E.; Cockburn, M.; Pfeifer, G.P. Wavelength Dependence of Ultraviolet Radiation-induced DNA Damage as Determined by Laser Irradiation Suggests That Cyclobutane Pyrimidine Dimers Are the Principal DNA Lesions Produced by Terrestrial Sunlight. FASEB J. 2011, 25, 3079–3091. [Google Scholar] [CrossRef]
  13. Buonanno, M.; Welch, D.; Shuryak, I.; Brenner, D.J. Far-UVC Light (222 Nm) Efficiently and Safely Inactivates Airborne Human Coronaviruses. Sci. Rep. 2020, 10, 10285. [Google Scholar] [CrossRef]
  14. Dai, T.; Vrahas, M.S.; Murray, C.K.; Hamblin, M.R. Ultraviolet C Irradiation: An Alternative Antimicrobial Approach to Localized Infections? Expert Rev. Anti Infect. Ther. 2012, 10, 185–195. [Google Scholar] [CrossRef] [PubMed]
  15. Beck, S.E.; Ryu, H.; Boczek, L.A.; Cashdollar, J.L.; Jeanis, K.M.; Rosenblum, J.S.; Lawal, O.R.; Linden, K.G. Evaluating UV-C LED Disinfection Performance and Investigating Potential Dual-Wavelength Synergy; Elsevier Ltd.: Amsterdam, The Netherlands, 2017; Volume 109, ISBN 3034924798. [Google Scholar]
  16. Brash, D.E. UV Signature Mutations. Photochem. Photobiol. 2015, 91, 15–26. [Google Scholar] [CrossRef]
  17. Sinha, R.P.; Häder, D.P. UV-Induced DNA Damage and Repair: A Review. Photochem. Photobiol. Sci. 2002, 1, 225–236. [Google Scholar] [CrossRef] [PubMed]
  18. Karu, T.I.; Pyatibrat, L.V.; Kolyakov, S.F.; Afanasyeva, N.I. Absorption Measurements of a Cell Monolayer Relevant to Phototherapy: Reduction of Cytochrome c Oxidase under near IR Radiation. J. Photochem. Photobiol. B 2005, 81, 98–106. [Google Scholar] [CrossRef]
  19. Zhao, C.S.; Shao, N.F.; Yang, S.T.; Ren, H.; Ge, Y.R.; Zhang, Z.S.; Feng, P.; Liu, W.L. Quantitative Assessment of the Effects of Human Activities on Phytoplankton Communities in Lakes and Reservoirs. Sci. Total Environ. 2019, 665, 213–225. [Google Scholar] [CrossRef]
  20. Oguma, K.; Katayama, H.; Ohgaki, S. Photoreactivation of Escherichia coli after Low- or Medium-Pressure UV Disinfection Determined by an Endonuclease Sensitive Site Assay. Appl. Environ. Microbiol. 2002, 68, 6029–6035. [Google Scholar] [CrossRef]
  21. Fan, X.; Huang, R.; Chen, H. Application of Ultraviolet C Technology for Surface Decontamination of Fresh Produce. Trends Food Sci. Technol. 2017, 70, 9–19. [Google Scholar] [CrossRef]
  22. Liltved, H.; Landfald, B. Effects of High Intensity Light on Ultraviolet-Irradiated and Non-Irradiated Fish Pathogenic Bacteria. Water Res. 2000, 34, 481–486. [Google Scholar] [CrossRef]
  23. Chevrefils, G.; Caron, É.; Wright, H.; Sakamoto, G. UV Dose Required to Achieve Incremental Log Inactivation of Bacteria, Protozoa and Viruses. IUVA News 2006, 8, 38–45. [Google Scholar]
  24. Martínez-Hernández, G.B.; Huertas, J.P.; Navarro-Rico, J.; Gómez, P.A.; Artés, F.; Palop, A.; Artés-Hernández, F. Inactivation Kinetics of Foodborne Pathogens by UV-C Radiation and Its Subsequent Growth in Fresh-Cut Kailan-Hybrid Broccoli. Food Microbiol. 2015, 46, 263–271. [Google Scholar] [CrossRef] [PubMed]
  25. Lim, W.; Harrison, M.A. Effectiveness of UV light as a means to reduce Salmonella contamination on tomatoes and food contact surfaces. Food Control 2016, 66, 166–173. [Google Scholar] [CrossRef]
  26. Liu, C.; Huang, Y.; Chen, H. Inactivation of Escherichia coli O157: H7 and Salmonella Enterica on Blueberries in Water Using Ultraviolet Light. J. Food Sci. 2015, 80, M1532–M1537. [Google Scholar] [CrossRef] [PubMed]
  27. Sommer, R.; Lhotsky, M.; Haider, T.; Cabaj, A. UV Inactivation, Liquid-Holding Recovery, and Photoreactivation of Escherichia coli O157 and Other Pathogenic Escherichia coli Strains in Water. J. Food Prot. 2000, 63, 1015–1020. [Google Scholar] [CrossRef]
  28. Chun, H.H.; Kim, J.Y.; Song, K. Bin Inactivation of Foodborne Pathogens in Ready-to-Eat Salad Using UV-C Irradiation. Food Sci. Biotechnol. 2010, 19, 547–551. [Google Scholar] [CrossRef]
  29. Allende, A.; McEvoy, J.L.; Luo, Y.; Artes, F.; Wang, C.Y. Effectiveness of Two-Sided UV-C Treatments in Inhibiting Natural Microflora and Extending the Shelf-Life of Minimally Processed “Red Oak Leaf” Lettuce. Food Microbiol. 2006, 23, 241–249. [Google Scholar] [CrossRef]
  30. Kim, Y.H.; Jeong, S.G.; Back, K.H.; Park, K.H.; Chung, M.S.; Kang, D.H. Effect of Various Conditions on Inactivation of Escherichia coli O157:H7, Salmonella typhimurium, and Listeria monocytogenes in Fresh-Cut Lettuce Using Ultraviolet Radiation. Int. J. Food Microbiol. 2013, 166, 349–355. [Google Scholar] [CrossRef]
  31. Chang, J.C.; Ossoff, S.F.; Lobe, D.C.; Dorfman, M.H.; Dumais, C.M.; Qualls, R.G.; Johnson, J.D. UV Inactivation of Pathogenic and Indicator Microorganisms. Appl. Environ. Microbiol. 1985, 49, 1361–1365. [Google Scholar] [CrossRef]
  32. Gryko, Ł.; Błaszczak, U.J.; Zajkowski, M. The Impact of Time and Temperature of Operation on the Characteristics of High-Power UVC LEDs and Their Disinfection Efficiency. Appl. Sci. 2023, 13, 12886. [Google Scholar] [CrossRef]
  33. Dai, T.; Tegos, G.P.; Rolz-Cruz, G.; Cumbie, W.E.; Hamblin, M.R. Ultraviolet C Inactivation of Dermatophytes: Implications for Treatment of Onychomycosis. Br. J. Dermatol. 2008, 158, 1239–1246. [Google Scholar] [CrossRef]
  34. Alcantara-Diaz, D. Divergent Adaptation of Escherichia coli to Cyclic Ultraviolet Light Exposures. Mutagenesis 2004, 19, 349–354. [Google Scholar] [CrossRef]
  35. DIN 5031-10:2018-03; Optical Radiation Physics and Illuminating Engineering—Part 10: Photobiologically Effective Radiation, Quantities, Symbols and Action Spectra. European Standard: Berlin, Germany, 2018.
  36. Madronich, S. The Atmosphere and UV-B Radiation at Ground Level. In Environmental UV Photobiology; Young, A.R., Moan, J., Björn, L.O., Nultsch, W., Eds.; Springer: Boston, MA, USA, 1993; pp. 1–39. ISBN 978-1-4899-2408-7. [Google Scholar]
  37. Sharma, V.K.; Demir, H.V. Bright Future of Deep-Ultraviolet Photonics: Emerging UVC Chip-Scale Light-Source Technology Platforms, Benchmarking, Challenges, and Outlook for UV Disinfection. ACS Photonics 2022, 9, 1513–1521. [Google Scholar] [CrossRef]
  38. Paul, J.; Kaneda, Y.; Wang, T.-L.; Lytle, C.; Moloney, J.V.; Jones, R.J. Doppler-Free Spectroscopy of Mercury at 2537 nm Using a High-Power, Frequency-Quadrupled, Optically Pumped External-Cavity Semiconductor Laser. Opt. Lett. 2011, 36, 61. [Google Scholar] [CrossRef] [PubMed]
  39. Obileke, K.; Onyeaka, H.; Miri, T.; Nwabor, O.F.; Hart, A.; Al-Sharify, Z.T.; Al-Najjar, S.; Anumudu, C. Recent Advances in Radio Frequency, Pulsed Light, and Cold Plasma Technologies for Food Safety. J. Food Process Eng. 2022, 45, e14138. [Google Scholar] [CrossRef]
  40. Bergman, R.S. Germicidal UV Sources and Systems. Photochem. Photobiol. 2021, 97, 466–470. [Google Scholar] [CrossRef] [PubMed]
  41. UN Environment. Minamata Convention on Mercury Text and Annexes, 2024 ed.; UN Environment: Nairobi, Kenya, 2024. [Google Scholar]
  42. Mildren, R.P.; Carman, R.J. Enhanced Performance of a Dielectric Barrier Discharge Lamp Using Short-Pulsed Excitation. J. Phys. D Appl. Phys. 2001, 34, L1–L6. [Google Scholar] [CrossRef]
  43. Kogelschatz, U. Excimer Lamps: History, Discharge Physics, and Industrial Applications; Tarasenko, V.F., Ed.; SPIE: Bellingham, WA, USA, 2004; pp. 272–286. [Google Scholar]
  44. Masoud, N.M.; Murnick, D.E. High Efficiency Fluorescent Excimer Lamps: An Alternative to Mercury Based UVC Lamps. Rev. Sci. Instrum. 2013, 84, 123108. [Google Scholar] [CrossRef]
  45. Susilo, N.; Hagedorn, S.; Jaeger, D.; Miyake, H.; Zeimer, U.; Reich, C.; Neuschulz, B.; Sulmoni, L.; Guttmann, M.; Mehnke, F.; et al. AlGaN-Based Deep UV LEDs Grown on Sputtered and High Temperature Annealed AlN/Sapphire. Appl. Phys. Lett. 2018, 112, 041110. [Google Scholar] [CrossRef]
  46. Kneissl, M.; Seong, T.-Y.; Han, J.; Amano, H. The Emergence and Prospects of Deep-Ultraviolet Light-Emitting Diode Technologies. Nat. Photonics 2019, 13, 233–244. [Google Scholar] [CrossRef]
  47. Takano, T.; Mino, T.; Sakai, J.; Noguchi, N.; Tsubaki, K.; Hirayama, H. Deep-Ultraviolet Light-Emitting Diodes with External Quantum Efficiency Higher than 20% at 275 Nm Achieved by Improving Light-Extraction Efficiency. Appl. Phys. Express 2017, 10, 031002. [Google Scholar] [CrossRef]
  48. Ivanov, S.V.; Jmerik, V.N.; Nechaev, D.V.; Kozlovsky, V.I.; Tiberi, M.D. E-beam Pumped Mid-UV Sources Based on MBE-grown AlGaN MQW. Phys. Status Solidi (A) 2015, 212, 1011–1016. [Google Scholar] [CrossRef]
  49. Kang, Y.; Zhao, J.; Wu, J.; Zhang, L.; Zhao, J.; Zhang, Y.; Zhao, Y.; Wang, X. Superior Deep-Ultraviolet Source Pumped by an Electron Beam for NLOS Communication. IEEE Trans. Electron. Devices 2020, 67, 3391–3394. [Google Scholar] [CrossRef]
  50. Li, D.; Jiang, K.; Sun, X.; Guo, C. AlGaN Photonics: Recent Advances in Materials and Ultraviolet Devices. Adv. Opt. Photonics 2018, 10, 43. [Google Scholar] [CrossRef]
  51. Watanabe, K.; Taniguchi, T.; Niiyama, T.; Miya, K.; Taniguchi, M. Far-Ultraviolet Plane-Emission Handheld Device Based on Hexagonal Boron Nitride. Nat. Photonics 2009, 3, 591–594. [Google Scholar] [CrossRef]
  52. Moffatt, J.E.; Tsiminis, G.; Klantsataya, E.; de Prinse, T.J.; Ottaway, D.; Spooner, N.A. A Practical Review of Shorter than Excitation Wavelength Light Emission Processes. Appl. Spectrosc. Rev. 2020, 55, 327–349. [Google Scholar] [CrossRef]
  53. He, G.S.; Tan, L.-S.; Zheng, Q.; Prasad, P.N. Multiphoton Absorbing Materials: Molecular Designs, Characterizations, and Applications. Chem. Rev. 2008, 108, 1245–1330. [Google Scholar] [CrossRef]
  54. Zhang, W.; Yang, S.; Li, J.; Gao, W.; Deng, Y.; Dong, W.; Zhao, C.; Lu, G. Visible-to-Ultraviolet Upconvertion: Energy Transfer, Material Matrix, and Synthesis Strategies. Appl. Catal. B 2017, 206, 89–103. [Google Scholar] [CrossRef]
  55. Auzel, F. Upconversion Processes in Coupled Ion Systems. J. Lumin. 1990, 45, 341–345. [Google Scholar] [CrossRef]
  56. van der Ende, B.M.; Aarts, L.; Meijerink, A. Lanthanide Ions as Spectral Converters for Solar Cells. Phys. Chem. Chem. Phys. 2009, 11, 11081. [Google Scholar] [CrossRef]
  57. Auzel, F. History of Upconversion Discovery and Its Evolution. J. Lumin. 2020, 223, 116900. [Google Scholar] [CrossRef]
  58. Auzel, F. Compteur Quantique Par Transfert d’energie Entre Deux Ions de Terres Rares Dans Un Tungstate Mixte et Dans Un Verre. CR Acad. Sci. Paris 1966, 262, 1016–1019. [Google Scholar]
  59. Dong, H.; Sun, L.-D.; Yan, C.-H. Energy Transfer in Lanthanide Upconversion Studies for Extended Optical Applications. Chem. Soc. Rev. 2015, 44, 1608–1634. [Google Scholar] [CrossRef]
  60. Fumes, A.C.; da Silva Telles, P.D.; Corona, S.A.M.; Borsatto, M.C. Effect of APDT on Streptococcus Mutans and Candida Albicans Present in the Dental Biofilm: Systematic Review. Photodiagnosis. Photodyn. Ther. 2018, 21, 363–366. [Google Scholar] [CrossRef]
  61. Youf, R.; Müller, M.; Balasini, A.; Thétiot, F.; Müller, M.; Hascoët, A.; Jonas, U.; Schönherr, H.; Lemercier, G.; Montier, T.; et al. Antimicrobial Photodynamic Therapy: Latest Developments with a Focus on Combinatory Strategies. Pharmaceutics 2021, 13, 1995. [Google Scholar] [CrossRef] [PubMed]
  62. Soukos, N.S.; Goodson, J.M. Photodynamic Therapy in the Control of Oral Biofilms. Periodontology 2000 2011, 55, 143–166. [Google Scholar] [CrossRef]
  63. Piksa, M.; Lian, C.; Samuel, I.C.; Pawlik, K.J.; Samuel, I.D.W.; Matczyszyn, K. The Role of the Light Source in Antimicrobial Photodynamic Therapy. Chem. Soc. Rev. 2023, 52, 1697–1722. [Google Scholar] [CrossRef] [PubMed]
  64. Mackay, A.M. The Evolution of Clinical Guidelines for Antimicrobial Photodynamic Therapy of Skin. Photochem. Photobiol. Sci. 2022, 21, 385–395. [Google Scholar] [CrossRef]
  65. Liu, W.; Sun, Y.; Zhou, B.; Chen, Y.; Liu, M.; Wang, L.; Qi, M.; Liu, B.; Dong, B. Near-Infrared Light Triggered Upconversion Nanocomposites with Multifunction of Enhanced Antimicrobial Photodynamic Therapy and Gas Therapy for Inflammation Regulation. J. Colloid Interface Sci. 2024, 663, 834–846. [Google Scholar] [CrossRef]
  66. Li, T.; Xue, C.; Wang, P.; Li, Y.; Wu, L. Photon Penetration Depth in Human Brain for Light Stimulation and Treatment: A Realistic Monte Carlo Simulation Study. J. Innov. Opt. Health Sci. 2017, 10, 1743002. [Google Scholar] [CrossRef]
  67. Hamblin, M.R. Upconversion in Photodynamic Therapy: Plumbing the Depths. Dalton Trans. 2018, 47, 8571–8580. [Google Scholar] [CrossRef]
  68. Liao, J.; Yang, L.; Wu, S.; Yang, Z.; Zhou, J.; Jin, D.; Guan, M. NIR-II Emissive Properties of 808 Nm-Excited Lanthanide-Doped Nanoparticles for Multiplexed in Vivo Imaging. J. Lumin. 2022, 242, 118597. [Google Scholar] [CrossRef]
  69. Yin, M.; Li, Z.; Ju, E.; Wang, Z.; Dong, K.; Ren, J.; Qu, X. Multifunctional Upconverting Nanoparticles for Near-Infrared Triggered and Synergistic Antibacterial Resistance Therapy. Chem. Commun. 2014, 50, 10488–10490. [Google Scholar] [CrossRef]
  70. Grüner, M.C.; Arai, M.S.; Carreira, M.; Inada, N.; de Camargo, A.S.S. Functionalizing the Mesoporous Silica Shell of Upconversion Nanoparticles To Enhance Bacterial Targeting and Killing via Photosensitizer-Induced Antimicrobial Photodynamic Therapy. ACS Appl. Bio Mater. 2018, 1, 1028–1036. [Google Scholar] [CrossRef] [PubMed]
  71. Dong, K.; Ju, E.; Gao, N.; Wang, Z.; Ren, J.; Qu, X. Synergistic Eradication of Antibiotic-Resistant Bacteria Based Biofilms in Vivo Using a NIR-Sensitive Nanoplatform. Chem. Commun. 2016, 52, 5312–5315. [Google Scholar] [CrossRef] [PubMed]
  72. Li, S.; Cui, S.; Yin, D.; Zhu, Q.; Ma, Y.; Qian, Z.; Gu, Y. Dual Antibacterial Activities of a Chitosan-Modified Upconversion Photodynamic Therapy System against Drug-Resistant Bacteria in Deep Tissue. Nanoscale 2017, 9, 3912–3924. [Google Scholar] [CrossRef]
  73. Liu, W.; Zhang, Y.; You, W.; Su, J.; Yu, S.; Dai, T.; Huang, Y.; Chen, X.; Song, X.; Chen, Z. Near-Infrared-Excited Upconversion Photodynamic Therapy of Extensively Drug-Resistant Acinetobacter baumannii Based on Lanthanide Nanoparticles. Nanoscale 2020, 12, 13948–13957. [Google Scholar] [CrossRef]
  74. Xu, F.; Zhao, Y.; Hu, M.; Zhang, P.; Kong, N.; Liu, R.; Liu, C.; Choi, S.K. Lanthanide-Doped Core–Shell Nanoparticles as a Multimodality Platform for Imaging and Photodynamic Therapy. Chem. Commun. 2018, 54, 9525–9528. [Google Scholar] [CrossRef]
  75. Xu, J.; Liu, N.; Wu, D.; Gao, Z.; Song, Y.-Y.; Schmuki, P. Upconversion Nanoparticle-Assisted Payload Delivery from TiO2 under Near-Infrared Light Irradiation for Bacterial Inactivation. ACS Nano 2020, 14, 337–346. [Google Scholar] [CrossRef]
  76. Nsubuga, A.; Morice, K.; Fayad, N.; Pini, F.; Josserand, V.; Le Guével, X.; Alhabi, A.; Henry, M.; Puchán Sánchez, D.; Plassais, N.; et al. Sub 20 nm Upconversion Photosensitizers for Near-Infrared Photodynamic Theranostics. Adv. Funct. Mater. 2025, 35, 2410077. [Google Scholar] [CrossRef]
  77. Karami, A.; Farivar, F.; de Prinse, T.J.; Rabiee, H.; Kidd, S.; Sumby, C.J.; Bi, J. Facile Multistep Synthesis of ZnO-Coated β-NaYF4:Yb/Tm Upconversion Nanoparticles as an Antimicrobial Photodynamic Therapy for Persistent Staphylococcus Aureus Small Colony Variants. ACS Appl. Bio Mater. 2021, 4, 6125–6136. [Google Scholar] [CrossRef]
  78. Wang, F.; Deng, R.; Wang, J.; Wang, Q.; Han, Y.; Zhu, H.; Chen, X.; Liu, X. Tuning Upconversion through Energy Migration in Core–Shell Nanoparticles. Nat. Mater. 2011, 10, 968–973. [Google Scholar] [CrossRef]
  79. Fu, X.; Fu, S.; Lu, Q.; Zhang, J.; Wan, P.; Liu, J.; Zhang, Y.; Chen, C.-H.; Li, W.; Wang, H.; et al. Excitation Energy Mediated Cross-Relaxation for Tunable Upconversion Luminescence from a Single Lanthanide Ion. Nat. Commun. 2022, 13, 4741. [Google Scholar] [CrossRef] [PubMed]
  80. Chen, X.; Jin, L.; Sun, T.; Kong, W.; Yu, S.F.; Wang, F. Energy Migration Upconversion in Ce(III)-Doped Heterogeneous Core−Shell−Shell Nanoparticles. Small 2017, 13, 1701479. [Google Scholar] [CrossRef]
  81. Su, Q.; Wei, H.-L.; Liu, Y.; Chen, C.; Guan, M.; Wang, S.; Su, Y.; Wang, H.; Chen, Z.; Jin, D. Six-Photon Upconverted Excitation Energy Lock-in for Ultraviolet-C Enhancement. Nat. Commun. 2021, 12, 4367. [Google Scholar] [CrossRef]
  82. Brik, M.G.; Ma, C.-G. Theoretical Spectroscopy of Transition Metal and Rare Earth Ions: From Free State to Crystal Field; Jenny Stanford Publishing: Singapore, 2020; ISBN 9780429278754. [Google Scholar]
  83. Carnall, W.T.; Goodman, G.L.; Rajnak, K.; Rana, R.S. A Systematic Analysis of the Spectra of the Lanthanides Doped into Single Crystal LaF3. J. Chem. Phys. 1989, 90, 3443–3457. [Google Scholar] [CrossRef]
  84. Sugar, J. Analysis of the Spectrum of Triply Ionized Praseodymium (Pr Iv). J. Opt. Soc. Am. 1965, 55, 1058. [Google Scholar] [CrossRef]
  85. Crosswhite, H.M.; Dieke, G.H.; Carter, W.J. Free-Ion and Crystalline Spectra of Pr3+ (Pr IV). J. Chem. Phys. 1965, 43, 2047–2054. [Google Scholar] [CrossRef]
  86. Kramida, A.; Ralchenko, Y.; Reader, J. NIST ASD Team NIST Atomic Spectra Database (Ver. 5.12). Available online: https://www.nist.gov/pml/atomic-spectra-database (accessed on 5 March 2025).
  87. Srivastava, A.M.; Jennings, M.; Collins, J. The Interconfigurational (4f15d1 → 4f2) Luminescence of Pr3+ in LuPO4, K3Lu(PO4)2 and LiLuSiO4. Opt. Mater. 2012, 34, 1347–1352. [Google Scholar] [CrossRef]
  88. Antić, Ž.; Racu, A.V.; Medić, M.; Alodhayb, A.N.; Kuzman, S.; Brik, M.G.; Dramićanin, M.D. Concentration and Temperature Dependence of Pr3+ F-f Emissions in La(PO3)3. Opt. Mater. 2024, 150, 115226. [Google Scholar] [CrossRef]
  89. Jüstel, T.; Mayr, W.; Schmidt, P.J.; Wiechert, D. On the Host Lattice Dependence of the 4fn−15d → 4fn Emission of Pr3+ and Nd3+. Available online: https://www.fh-muenster.de/ciw/downloads/personal/juestel/juestel/On_the_Host_Lattice_Dependence_of_the_4fn-15d1_Emission_of_Pr3__and_Nd3___November_2001_.pdf (accessed on 5 March 2025).
  90. Dorenbos, P. The 5d Level Positions of the Trivalent Lanthanides in Inorganic Compounds. J. Lumin. 2000, 91, 155–176. [Google Scholar] [CrossRef]
  91. Dorenbos, P. The 4f ↔ 4f − 15d Transitions of the Trivalent Lanthanides in Halogenides and Chalcogenides. J. Lumin. 2000, 91, 91–106. [Google Scholar] [CrossRef]
  92. Laroche, M.; Bettinelli, M.; Girard, S.; Moncorgé, R. F–d Luminescence of Pr3+ and Ce3+ in the Chloro-Elpasolite Cs2NaYCl6. Chem. Phys. Lett. 1999, 311, 167–172. [Google Scholar] [CrossRef]
  93. Tanner, P.A. Spectra, Energy Levels and Energy Transfer in High Symmetry Lanthanide Compounds. In Transition Metal and Rare Earth Compounds III Excited States, Transitions, Interactions; Yersin, H., Ed.; Springer: Berlin/Heidelberg, Germany, 2004; Volume 241, pp. 167–278. [Google Scholar]
  94. Li, Y.; Lu, H.; Li, J.; Miao, X. Luminescence of Several Rare Earth Ions in LaOI. J. Lumin. 1986, 35, 107–109. [Google Scholar] [CrossRef]
  95. Wang, D.; Guo, Y.; Zhang, E.; Chao, X.; Yu, L.; Luo, J.; Zhang, W.; Yin, M. Synthesis and NIR-to-Violet, Blue, Green, Red Upconversion Fluorescence of Er3+:LaOBr. J. Alloys Compd. 2005, 397, 1–4. [Google Scholar] [CrossRef]
  96. Rebrova, N.; Grippa, A.; Zdeb, P.; Dereń, P.J. Blue to UV Upconversion Properties of Pr3+ Doped ACaF3 (A = K, Rb, Cs) Phosphors. Scr. Mater. 2025, 255, 116395. [Google Scholar] [CrossRef]
  97. Daniel, P.; Rousseau, M.; Toulouse, J. Raman Scattering Study of Potassium Calcium Fluoride KCaF3. Phys. Status Solidi (B) 1997, 203, 327–335. [Google Scholar] [CrossRef]
  98. Ridou, C.; Rousseau, M.; Gervais, F. The Temperature Dependence of the Infrared Reflection Spectra in the Fluoperovskites RbCaF3, CsCaF3 and KZnF3. J. Phys. C Solid State Phys. 1986, 19, 5757–5767. [Google Scholar] [CrossRef]
  99. Sarantopoulou, E.; Abdulsabirov, R.Y.; Korableva, S.L.; Cefalas, A.C.; Dubinskii, M.A.; Naumov, A.K.; Semashko, V.V.; Nicolaides, C.A. VUV and UV Fluorescence and Absorption Studies of Pr3+-Doped LiLuF4 Single Crystals. Opt. Lett. 1994, 19, 499. [Google Scholar] [CrossRef]
  100. Ma, Y.; Wen, T.; Liu, K.; Jiang, D.; Zhao, M.-H.; Lin, C.; Wang, Y. Pressure-Induced Structural Phase Transition, Irreversible Amorphization and Upconversion Luminescence Enhancement in Ln3+-Codoped LiYF4 and LiLuF4. J. Mater. Chem. C Mater. 2023, 11, 6588–6596. [Google Scholar] [CrossRef]
  101. Vink, A.P.; van der Kolk, E.; Dorenbos, P.; van Eijk, C.W.E. Opposite Parity 4f−15d1 States of Ce3+ and Pr3+ in MSO4 (M = Ca, Sr, Ba). J. Alloys Compd. 2002, 341, 338–341. [Google Scholar] [CrossRef]
  102. Smith, D.H.; Seshadri, K.S. Infrared Spectra of Mg2Ca(SO4)3, MgSO4, Hexagonal CaSO4, and Orthorhombic CaSO4. Spectrochim. Acta A Mol. Biomol. Spectrosc. 1999, 55, 795–805. [Google Scholar] [CrossRef]
  103. Yang, Y.-M.; Li, Z.-Y.; Zhang, J.-Y.; Lu, Y.; Guo, S.-Q.; Zhao, Q.; Wang, X.; Yong, Z.-J.; Li, H.; Ma, J.-P.; et al. X-Ray-Activated Long Persistent Phosphors Featuring Strong UVC Afterglow Emissions. Light Sci. Appl. 2018, 7, 88. [Google Scholar] [CrossRef] [PubMed]
  104. Nair, R.G.; Nigam, S.; Sudarsan, V.; Vatsa, R.K.; Jain, V.K. YBO3 versus Y3BO6 Host on Tb3+ Luminescence. J. Lumin. 2018, 195, 271–277. [Google Scholar] [CrossRef]
  105. Chen, W.; Li, L.; Liang, H.; Tian, Z.; Su, Q.; Zhang, G. Luminescence of Pr3+ in La2CaB10O19: Simultaneous Observation PCE and f–d Emission in a Single Host. Opt. Mater. 2009, 32, 115–120. [Google Scholar] [CrossRef]
  106. Szymborska-Małek, K.; Ptak, M.; Tomaszewski, P.E.; Majchrowski, A. Raman and IR Spectroscopic Study of a Nonlinear Optical Crystal, La2CaB10O19. Vib. Spectrosc. 2016, 82, 53–59. [Google Scholar] [CrossRef]
  107. Zeler, J.; Sulollari, M.; Meijerink, A.; Bettinelli, M.; Zych, E. Chemical Stabilization of Eu2+ in LuPO4 and YPO4 Hosts and Its Peculiar Sharp Line Luminescence. J. Alloys Compd. 2020, 844, 156096. [Google Scholar] [CrossRef]
  108. Kappelhoff, J.; Keil, J.-N.; Kirm, M.; Makhov, V.N.; Chernenko, K.; Möller, S.; Jüstel, T. Spectroscopic Studies on Pr3+ Doped YPO4 and LuPO4 upon Vacuum Ultraviolet (VUV) and Synchrotron Radiation Excitation. Chem. Phys. 2022, 562, 111646. [Google Scholar] [CrossRef]
  109. Keil, J.-N.; Jenneboer, H.; Jüstel, T. Temperature Dependent Luminescence of Pr3+ Doped NaCaPO4. J. Lumin. 2021, 238, 118307. [Google Scholar] [CrossRef]
  110. Jastrzębski, W.; Sitarz, M.; Rokita, M.; Bułat, K. Infrared Spectroscopy of Different Phosphates Structures. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2011, 79, 722–727. [Google Scholar] [CrossRef]
  111. Ivanovskikh, K.V.; Pustovarov, V.A.; Omelkov, S.; Kirm, M.; Piccinelli, F.; Bettinelli, M. Phase Transition, Radio- and Photoluminescence of K3Lu(PO4)2 Doped with Pr3+ Ions. J. Lumin. 2021, 230, 117749. [Google Scholar] [CrossRef]
  112. Pelczarska, A.; Watras, A.; Godlewska, P.; Radomińska, E.; Macalik, L.; Szczygieł, I.; Hanuza, J.; Dereń, P.J. Structural, Raman, FT-IR and Optical Properties of RbY2(PO4)3 and Rb3La(PO4)2 Doped with Eu3+ Ions. New J. Chem. 2015, 39, 8474–8483. [Google Scholar] [CrossRef]
  113. Keil, J.-N.; Lindfeld, E.; Jüstel, T. Synthesis and Characterization of Sr3(PO4)2:Pr3+, Si4+. J. Lumin. 2020, 225, 117376. [Google Scholar] [CrossRef]
  114. Zhai, S.; Lin, C.-C.; Xue, W. Raman Spectra of Sr3(PO4)2 and Ba3(PO4)2 Orthophosphates at Various Temperatures. Vib. Spectrosc. 2014, 70, 6–11. [Google Scholar] [CrossRef]
  115. Rebrova, N.; Lisiecki, R.; Zdeb-Stańczykowska, P.; Zorenko, Y.; Voloshinovskii, A.; Pushak, A.; Dereń, P.J. Optical and Upconversion Properties of AY(PO4)3:Pr3+ (A = Sr, Ba) Phosphors. J. Phys. Chem. C 2025, 129, 1873–1884. [Google Scholar] [CrossRef]
  116. Guan, A.; Chen, P.; Zhou, L.; Wang, G.; Zhang, X.; Tang, J. Color-Tunable Emission and Energy Transfer Investigation in Sr3Y(PO4)3:Ce3+, Tb3+ Phosphors for White LEDs. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2017, 173, 53–58. [Google Scholar] [CrossRef] [PubMed]
  117. Li, K.; Shang, M.; Lian, H.; Lin, J. Recent Development in Phosphors with Different Emitting Colors via Energy Transfer. J. Mater. Chem. C Mater. 2016, 4, 5507–5530. [Google Scholar] [CrossRef]
  118. Lemański, K.; Bezkrovna, O.; Rebrova, N.; Lisiecki, R.; Zdeb, P.; Dereń, P.J. UVC Stokes and Anti-Stokes Emission of Ca9Y(PO4)7 Polycrystals Doped with Pr3+ Ions. Molecules 2024, 29, 2084. [Google Scholar] [CrossRef]
  119. Su, C.; Ao, L.; Zhang, Z.; Zhai, Y.; Chen, J.; Tang, Y.; Liu, L.; Fang, L. Crystal Structure, Raman Spectra and Microwave Dielectric Properties of Novel Temperature-Stable LiYbSiO4 Ceramics. Ceram. Int. 2020, 46, 19996–20003. [Google Scholar] [CrossRef]
  120. Yin, Z.; Yuan, P.; Zhu, Z.; Li, T.; Yang, Y. Pr3+ Doped Li2SrSiO4: An Efficient Visible-Ultraviolet C up-Conversion Phosphor. Ceram. Int. 2021, 47, 4858–4863. [Google Scholar] [CrossRef]
  121. Wu, H.; Zhang, B.; Yu, H.; Hu, Z.; Wang, J.; Wu, Y.; Halasyamani, P.S. Designing Silicates as Deep-UV Nonlinear Optical (NLO) Materials Using Edge-Sharing Tetrahedra. Angew. Chem. 2020, 132, 9007–9011. [Google Scholar] [CrossRef]
  122. Rebrova, N.; Zdeb, P.; Dereń, P.J. Synthesis and Upconversion Luminescence of LiY(SiO)6O2 Phosphor Doped with Pr3+. J. Phys. Chem. C 2024, 128, 9090–9098. [Google Scholar] [CrossRef]
  123. Li, Y.; Liang, D.; Zhang, X.; Xiong, Z.; Tang, B.; Si, F.; Fang, Z.; Shi, Z.; Chen, J.; Wang, F.; et al. Sintering Behavior, Crystal Structure, and Microwave Dielectric Properties of a Novel NaY9Si6O26 Ceramic. J. Am. Ceram. Soc. 2024, 107, 4077–4085. [Google Scholar] [CrossRef]
  124. Yan, S.; Liang, Y.; Chen, Y.; Liu, J.; Chen, D.; Pan, Z. Ultraviolet-C Persistent Luminescence from the Lu2SiO5:Pr3+ Persistent Phosphor for Solar-Blind Optical Tagging. Dalton Trans. 2021, 50, 8457–8466. [Google Scholar] [CrossRef]
  125. Han, L.; Song, F.; Chen, S.-Q.; Zou, C.-G.; Yu, X.-C.; Tian, J.-G.; Xu, J.; Xu, X.; Zhao, G. Intense Upconversion and Infrared Emissions in Er3+–Yb3+ Codoped Lu2SiO5 and (Lu0.5Gd0.5)2SiO5 Crystals. Appl. Phys. Lett. 2008, 93, 011110. [Google Scholar] [CrossRef]
  126. Cates, E.L.; Wilkinson, A.P.; Kim, J.-H. Delineating Mechanisms of Upconversion Enhancement by Li+ Codoping in Y2SiO5:Pr3+. J. Phys. Chem. C 2012, 116, 12772–12778. [Google Scholar] [CrossRef]
  127. Denoyer, A.; Lévesque, Y.; Jandl, S.; Guillot-Noël, O.; Goldner, P.; Viana, B.; Thibault, F.; Pelenc, D. Crystal Field Study of Ytterbium Doped Lu2SiO5 and Y2SiO5 under a Magnetic Field. J. Phys. Condens. Matter 2008, 20, 125227. [Google Scholar] [CrossRef]
  128. Auzel, F. Upconversion and Anti-Stokes Processes with f and d Ions in Solids. Chem. Rev. 2004, 104, 139–174. [Google Scholar] [CrossRef]
  129. Remillieux, A.; Jacquier, B. IR-to-Visible up-Conversion Mechanisms in Pr3+-Doped ZBLAN Fluoride Glasses and Fibers. J. Lumin. 1996, 68, 279–289. [Google Scholar] [CrossRef]
  130. Balda, R.; Fernández, J.; Saéz de Ocáriz, I.; Voda, M.; García, A.J.; Khaidukov, N. Laser Spectroscopy of Pr3+ Ions in LiKY1−xPrxF5 Single Crystals. Phys. Rev. B 1999, 59, 9972–9980. [Google Scholar] [CrossRef]
  131. Dereń, P.J.; Mahiou, R.; Stręk, W.; Bednarkiewicz, A.; Bertrand, G. Up-Conversion in KYb(WO4)2:Pr3+ Crystal. Opt. Mater. 2002, 19, 145–148. [Google Scholar] [CrossRef]
  132. Schröder, F.; Pues, P.; Enseling, D.; Jüstel, T. On the Quantum Yield Determination of UV Emitting Up-Converters. Luminescence 2023, 38, 702–708. [Google Scholar] [CrossRef] [PubMed]
  133. Zalucha, D.J.; Wright, J.C.; Fong, F.K. Energy Transfer Upconversion in LaF3:Pr3+. J. Chem. Phys. 1973, 59, 997–1001. [Google Scholar] [CrossRef]
  134. Zalucha, D.J.; Sell, J.A.; Fong, F.K. Infrared and Visible Photon Upconversion in LaCl3:Pr3+ (Nd3+). J. Chem. Phys. 1974, 60, 1660–1667. [Google Scholar] [CrossRef]
  135. Auzel, F.E. Materials and Devices Using Double-Pumped-Phosphors with Energy Transfer. Proc. IEEE 1973, 61, 758–786. [Google Scholar] [CrossRef]
  136. Brown, M.R.; Whiting, J.S.S.; Shand, W.A. Ion—Ion Interactions in Rare-Earth-Doped LaF3. J. Chem. Phys. 1965, 43, 1–9. [Google Scholar] [CrossRef]
  137. Wright, J.C.; Zalucha, D.J.; Lauer, H.V.; Cox, D.E.; Fong, F.K. Laser Optical Double Resonance and Efficient Infrared Quantum Counter Upconversion in LaCl3:Pr3+ and LaF3:Pr3+. J. Appl. Phys. 1973, 44, 781–786. [Google Scholar] [CrossRef]
  138. Sun, C.L.; Li, J.F.; Hu, C.H.; Jiang, H.M.; Jiang, Z.K. Ultraviolet Upconversion in Pr3+:Y2SiO5 Crystal by Ar+ Laser (488 nm) Excitation. Eur. Phys. J. D 2006, 39, 303–306. [Google Scholar] [CrossRef]
  139. Jones, C.M.S.; Gakamsky, A.; Marques-Hueso, J. The Upconversion Quantum Yield (UCQY): A Review to Standardize the Measurement Methodology, Improve Comparability, and Define Efficiency Standards. Sci. Technol. Adv. Mater. 2021, 22, 810–848. [Google Scholar] [CrossRef]
  140. Kaiser, M.; Würth, C.; Kraft, M.; Hyppänen, I.; Soukka, T.; Resch-Genger, U. Power-Dependent Upconversion Quantum Yield of NaYF4:Yb3+,Er3+ Nano- and Micrometer-Sized Particles—Measurements and Simulations. Nanoscale 2017, 9, 10051–10058. [Google Scholar] [CrossRef]
  141. Homann, C.; Krukewitt, L.; Frenzel, F.; Grauel, B.; Würth, C.; Resch-Genger, U.; Haase, M. NaYF4:Yb, Er/NaYF4 Core/Shell Nanocrystals with High Upconversion Luminescence Quantum Yield. Angew. Chem. Int. Ed. 2018, 57, 8765–8769. [Google Scholar] [CrossRef]
  142. Cates, E.L.; Cho, M.; Kim, J.-H. Converting Visible Light into UVC: Microbial Inactivation by Pr3+-Activated Upconversion Materials. Environ. Sci. Technol. 2011, 45, 3680–3686. [Google Scholar] [CrossRef] [PubMed]
  143. Cates, E.L.; Li, F. Balancing Intermediate State Decay Rates for Efficient Pr3+ Visible-to-UVC Upconversion: The Case of β-Y2Si2O7:Pr3+. RSC Adv. 2016, 6, 22791–22796. [Google Scholar] [CrossRef]
  144. Du, Y.; Jin, Z.; Li, Z.; Sun, T.; Meng, H.; Jiang, X.; Wang, Y.; Peng, D.; Li, J.; Wang, A.; et al. Tuning the 5d State of Pr3+ in Oxyhalides for Efficient Deep Ultraviolet Upconversion. Adv. Opt. Mater. 2024, 12, 2400971. [Google Scholar] [CrossRef]
  145. Zdeb, P.; Rebrova, N.; Dereń, P.J. Discovering the Potential of High Phonon Energy Hosts in the Field of Visible-to-Ultraviolet C Upconversion. J. Phys. Chem. Lett. 2024, 15, 9356–9360. [Google Scholar] [CrossRef]
  146. Zhang, L.; Dong, G.; Peng, M.; Qiu, J. Comparative Investigation on the Spectroscopic Properties of Pr3+-Doped Boro-Phosphate, Boro-Germo-Silicate and Tellurite Glasses. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2012, 93, 223–227. [Google Scholar] [CrossRef]
  147. Naresh, V.; Ham, B.S. Influence of Multiphonon and Cross Relaxations on 3P0 and 1D2 Emission Levels of Pr3+ Doped Borosilicate Glasses for Broad Band Signal Amplification. J. Alloys Compd. 2016, 664, 321–330. [Google Scholar] [CrossRef]
  148. van Dijk, J.M.F.; Schuurmans, M.F.H. On the Nonradiative and Radiative Decay Rates and a Modified Exponential Energy Gap Law for 4 f–4 f Transitions in Rare-Earth Ions. J. Chem. Phys. 1983, 78, 5317–5323. [Google Scholar] [CrossRef]
  149. Kumar, M.V.V.; Gopal, K.R.; Reddy, R.R.; Reddy, G.V.L.; Hussain, N.S.; Jamalaiah, B.C. Application of Modified Judd–Ofelt Theory and the Evaluation of Radiative Properties of Pr3+-Doped Lead Telluroborate Glasses for Laser Applications. J. Non. Cryst. Solids 2013, 364, 20–27. [Google Scholar] [CrossRef]
  150. Malyukin, Y.V.; Masalov, A.A.; Zhmurin, P.N.; Znamenskii, N.V.; Petrenko, E.A.; Yukina, T.G. Two Mechanisms of 1D2 Fluorescence Quenching of Pr3+-doped Y2SiO5 Crystal. Phys. Status Solidi (B) 2003, 240, 655–662. [Google Scholar] [CrossRef]
  151. Wen, S.; Zhou, J.; Schuck, P.J.; Suh, Y.D.; Schmidt, T.W.; Jin, D. Future and Challenges for Hybrid Upconversion Nanosystems. Nat. Photonics 2019, 13, 828–838. [Google Scholar] [CrossRef]
  152. Zhao, H.; Zhang, L.; Lu, J.; Chai, S.; Wei, J.; Yu, Y.; Miao, R.; Zhong, L. Visible-UVC Upconversion Polymer Films for Prevention of Microbial Infection. J. Mater. Chem. B 2023, 11, 2745–2753. [Google Scholar] [CrossRef] [PubMed]
  153. Zhang, Y.; Luo, Y.; Fu, S.; Lv, X.; He, Q.; Ji, F.; Xu, X. Visible-to-UVC Driven Upconversion Photocatalyst Sterilization Efficiency and Mechanisms of β-NaYF4: Pr3+, Li+@BiOCl with a Core-Shell Structure. J. Environ. Manag. 2021, 288, 112394. [Google Scholar] [CrossRef] [PubMed]
  154. Cates, E.L.; Wilkinson, A.P.; Kim, J.H. Visible-to-UVC Upconversion Efficiency and Mechanisms of Lu7O6F9:Pr3+ and Y2SiO5:Pr3+ Ceramics. J. Lumin. 2015, 160, 202–209. [Google Scholar] [CrossRef]
  155. Tsang, M.Y.; Fałat, P.; Antoniak, M.A.; Ziniuk, R.; Zelewski, S.J.; Samoć, M.; Nyk, M.; Qu, J.; Ohulchanskyy, T.Y.; Wawrzyńczyk, D. Pr3+ Doped NaYF4 and LiYF4 Nanocrystals Combining Visible-to-UVC Upconversion and NIR-to-NIR-II Downconversion Luminescence Emissions for Biomedical Applications. Nanoscale 2022, 14, 14770–14778. [Google Scholar] [CrossRef]
  156. Lv, P.; Li, L.; Yin, Z.; Wang, C.; Yang, Y. Visible-to-Ultraviolet-C Upconverted Photon for Multifunction via Ca2SiO4:Pr3+. Opt. Lett. 2022, 47, 4435. [Google Scholar] [CrossRef]
  157. Fałat, P.; Tsang, M.Y.; Maliszewska, I.; Zelewski, S.J.; Cichy, B.; Ohulchanskyy, T.Y.; Samoć, M.; Nyk, M.; Wawrzyńczyk, D. Enhanced Biocidal Activity of Pr3+ Doped Yttrium Silicates by Tm3+ and Yb3+ Co-Doping. Mater. Adv. 2023, 4, 5827–5837. [Google Scholar] [CrossRef]
  158. Zhou, X.; Qiao, J.; Zhao, Y.; Han, K.; Xia, Z. Multi-Responsive Deep-Ultraviolet Emission in Praseodymium-Doped Phosphors for Microbial Sterilization. Sci. China Mater. 2022, 65, 1103–1111. [Google Scholar] [CrossRef]
  159. Zi, L.; Li, L.; Wang, C.; Yang, F.; Feng, S.; Lv, P.; Yang, Y. Triple-Responsive Visible-To-Ultraviolet-C Upconverted Photons for Multifunctional Applications. Adv. Opt. Mater. 2024, 12, 2301881. [Google Scholar] [CrossRef]
  160. Zhang, Y.; Yan, S.; Xiao, F.; Shan, X.; Lv, X.; Wang, W.; Liang, Y. Long-Persistent Far-UVC Light Emission in Pr3+-Doped Sr2P2O7 Phosphor for Microbial Sterilization. Inorg. Chem. Front. 2023, 10, 5958–5968. [Google Scholar] [CrossRef]
  161. Wang, C.; Tang, Y.; Pu, G.; Chen, W.; Deng, M.; Wang, J. Realizing Golden Ultraviolet C Emission of 265 nm by Oxygen Vacancies Engineering for 100% Sterilization Efficiency. Ceram. Int. 2024, 50, 30579–30586. [Google Scholar] [CrossRef]
  162. Rim, K.-T. Effects of Rare Earth Elements on the Environment and Human Health: A Literature Review. Toxicol. Environ. Health Sci. 2016, 8, 189–200. [Google Scholar] [CrossRef]
  163. Constantin, M.; Chifiriuc, M.C.; Vrancianu, C.O.; Petrescu, L.; Cristian, R.-E.; Crunteanu, I.; Grigore, G.A.; Chioncel, M.F. Insights into the Effects of Lanthanides on Mammalian Systems and Potential Applications. Environ. Res. 2024, 263, 120235. [Google Scholar] [CrossRef] [PubMed]
  164. Gnach, A.; Lipinski, T.; Bednarkiewicz, A.; Rybka, J.; Capobianco, J.A. Upconverting Nanoparticles: Assessing the Toxicity. Chem. Soc. Rev. 2015, 44, 1561–1584. [Google Scholar] [CrossRef]
  165. Buzea, C.; Pacheco, I.I.; Robbie, K. Nanomaterials and Nanoparticles: Sources and Toxicity. Biointerphases 2007, 2, MR17–MR71. [Google Scholar] [CrossRef] [PubMed]
  166. Yang, L.; Watts, D.J. Particle Surface Characteristics May Play an Important Role in Phytotoxicity of Alumina Nanoparticles. Toxicol. Lett. 2005, 158, 122–132. [Google Scholar] [CrossRef]
  167. Donaldson, K.; Brown, D.; Clouter, A.; Duffin, R.; MacNee, W.; Renwick, L.; Tran, L.; Stone, V. The Pulmonary Toxicology of Ultrafine Particles. J. Aerosol. Med. 2002, 15, 213–220. [Google Scholar] [CrossRef]
  168. Bahadar, H.; Maqbool, F.; Niaz, K.; Abdollahi, M. Toxicity of Nanoparticles and an Overview of Current Experimental Models. Iran Biomed. J. 2016, 20, 1–11. [Google Scholar] [CrossRef]
  169. Lovreglio, P.; D’Errico, M.N.; De Pasquale, P.; Gilberti, M.E.; Drago, I.; Panuzzo, L.; Lepera, A.; Serra, R.; Ferrara, F.; Basso, A.; et al. Environmental Factors Affecting the Urinary Excretion of Inorganic Arsenic in the General Population. Med. Lav. 2012, 103, 372–381. [Google Scholar]
  170. Selvaraj, V.; Bodapati, S.; Murray, E.; Rice, K.M.; Winston, N.; Shokuhfar, T.; Zhao, Y.; Blough, E. Cytotoxicity and Genotoxicity Caused by Yttrium Oxide Nanoparticles in HEK293 Cells. Int. J. Nanomed. 2014, 9, 1379–1391. [Google Scholar] [CrossRef]
  171. Gojova, A.; Guo, B.; Kota, R.S.; Rutledge, J.C.; Kennedy, I.M.; Barakat, A.I. Induction of Inflammation in Vascular Endothelial Cells by Metal Oxide Nanoparticles: Effect of Particle Composition. Environ. Health Perspect 2007, 115, 403–409. [Google Scholar] [CrossRef]
  172. Liu, X.; Xiang, Q.; Zhang, L.; Li, J.; Wu, Y. Occurrence of Rare Earth Elements in Umbilical Cord Serum and Association with Thyroid Hormones and Birth Outcomes in Newborns. Chemosphere 2024, 359, 142321. [Google Scholar] [CrossRef] [PubMed]
  173. Liu, Z.; Ding, Y.; Xie, S.; Hu, Y.; Xiao, H.; Liu, X.; Fan, X. Chronic Exposure to Yttrium Induced Cell Apoptosis in the Testis by Mediating Ca2+/IP3R1/CaMKII Signaling. Front. Public Health 2023, 11, 1104195. [Google Scholar] [CrossRef]
  174. Turyanskaya, A.; Rauwolf, M.; Pichler, V.; Simon, R.; Burghammer, M.; Fox, O.J.L.; Sawhney, K.; Hofstaetter, J.G.; Roschger, A.; Roschger, P.; et al. Detection and Imaging of Gadolinium Accumulation in Human Bone Tissue by Micro- and Submicro-XRF. Sci. Rep. 2020, 10, 6301. [Google Scholar] [CrossRef]
  175. Ergün, I.; Keven, K.; Uruç, I.; Ekmekçi, Y.; Canbakan, B.; Erden, I.; Karatan, O. The Safety of Gadolinium in Patients with Stage 3 and 4 Renal Failure. Nephrol. Dial. Transplant. 2006, 21, 697–700. [Google Scholar] [CrossRef]
  176. Grobner, T. Gadolinium—A Specific Trigger for the Development of Nephrogenic Fibrosing Dermopathy and Nephrogenic Systemic Fibrosis? Nephrol. Dial. Transplant. 2006, 21, 1104–1108. [Google Scholar] [CrossRef] [PubMed]
  177. Rydahl, C.; Thomsen, H.S.; Marckmann, P. High Prevalence of Nephrogenic Systemic Fibrosis in Chronic Renal Failure Patients Exposed to Gadodiamide, a Gadolinium-Containing Magnetic Resonance Contrast Agent. Investig. Radiol. 2008, 43, 141–144. [Google Scholar] [CrossRef]
  178. Akhtar, M.J.; Ahamed, M.; Alhadlaq, H. Gadolinium Oxide Nanoparticles Induce Toxicity in Human Endothelial Huvecs via Lipid Peroxidation, Mitochondrial Dysfunction and Autophagy Modulation. Nanomaterials 2020, 10, 1675. [Google Scholar] [CrossRef]
  179. Lattanzio, S.M.; Imbesi, F. Fibromyalgia Associated with Repeated Gadolinium Contrast-Enhanced MRI Examinations. Radiol. Case Rep. 2020, 15, 534–541. [Google Scholar] [CrossRef]
  180. Bilgin, B.; Adam, M.; Hekim, M.G.; Bulut, F.; Ozcan, M. Gadolinium-Based Contrast Agents Aggravate Mechanical and Thermal Hyperalgesia in a Nitroglycerine-Induced Migraine Model in Male Mice. Magn. Reson. Imaging 2024, 111, 67–73. [Google Scholar] [CrossRef]
  181. Edward, M.; Quinn, J.A.; Burden, A.D.; Newton, B.B.; Jardine, A.G. Effect of Different Classes of Gadolinium-Based Contrast Agents on Control and Nephrogenic Systemic Fibrosis-Derived Fibroblast Proliferation. Radiology 2010, 256, 735–743. [Google Scholar] [CrossRef]
  182. Becker, P.C.N.; Olsson, A.; Simpson, J.R. Erbium-Doped Fiber Amplifiers: Fundamentals and technology; Academic Press: San Diego, CA, USA, 1999; Volume 102, ISBN 0120845903. [Google Scholar]
  183. Cieślik, I.; Płocińska, M.; Płociński, T.; Zdunek, J.; Woźniak, M.J.; Bil, M.; Hirano, S. Influence of Polymeric Precursors on the Viability of Human Cells of Yttrium Aluminum Borates Nanoparticles Doped with Ytterbium Ions. Appl. Surf. Sci. 2019, 488, 874–886. [Google Scholar] [CrossRef]
  184. Antinori, S.; Versaci, C.; Fuhrberg, P.; Panci, C.; Caffa, B.; Gholami, G.H. Andrology: Seventeen Live Births after the Use of an Erbium-Yytrium Aluminium Garnet Laser in the Treatment of Male Factor Infertility. Hum. Reprod. 1994, 9, 1891–1896. [Google Scholar] [CrossRef] [PubMed]
  185. Feyerabend, F.; Fischer, J.; Holtz, J.; Witte, F.; Willumeit, R.; Drücker, H.; Vogt, C.; Hort, N. Evaluation of Short-Term Effects of Rare Earth and Other Elements Used in Magnesium Alloys on Primary Cells and Cell Lines. Acta Biomater. 2010, 6, 1834–1842. [Google Scholar] [CrossRef]
  186. Weltje, L.; Verhoof, L.R.C.W.; Verweij, W.; Hamers, T. Lutetium Speciation and Toxicity in a Microbial Bioassay: Testing the Free-Ion Model for Lanthanides. Environ. Sci. Technol. 2004, 38, 6597–6604. [Google Scholar] [CrossRef]
  187. Dash, A.; Pillai, M.R.A.; Knapp, F.F. Production of 177Lu for Targeted Radionuclide Therapy: Available Options. Nucl. Med. Mol. Imaging 2015, 49, 85–107. [Google Scholar] [CrossRef]
  188. Dolgikh, O.V.; Alekseev, V.B.; Dianova, D.G.; Vdovina, N.A. Features of the Immune Profile of Workers of a Non-Ferrous Metallurgy Enterprise in Conditions of Contamination of Biological Media with Rare Earth Elements (Using the Example of Thulium). Russ. J. Occup. Heralth Indust. Ecol. 2024, 64, 525–530. [Google Scholar] [CrossRef]
  189. Perry, J.; Minaei, E.; Engels, E.; Ashford, B.G.; McAlary, L.; Clark, J.R.; Gupta, R.; Tehei, M.; Corde, S.; Carolan, M.; et al. Thulium Oxide Nanoparticles as Radioenhancers for the Treatment of Metastatic Cutaneous Squamous Cell Carcinoma. Phys. Med. Biol. 2020, 65, 215018. [Google Scholar] [CrossRef]
  190. Genchi, G.; Carocci, A.; Lauria, G.; Sinicropi, M.S.; Catalano, A. Thallium Use, Toxicity, and Detoxification Therapy: An Overview. Appl. Sci. 2021, 11, 8322. [Google Scholar] [CrossRef]
  191. Singh, A.; Raj, A.; Shah, P.; Agrawal, N. Nanoparticles: An Activator of Oxidative Stress. In Toxicology of Nanoparticles: Insights from Drosophila; Agrawal, N., Shah, P., Eds.; Springer: Singapore, 2020; pp. 89–106. ISBN 978-981-15-5522-0. [Google Scholar]
  192. Koedrith, P.; Boonprasert, R.; Kwon, J.Y.; Kim, I.-S.; Seo, Y.R. Recent Toxicological Investigations of Metal or Metal Oxide Nanoparticles in Mammalian Models in Vitro and in Vivo: DNA Damaging Potential, and Relevant Physicochemical Characteristics. Mol. Cell Toxicol. 2014, 10, 107–126. [Google Scholar] [CrossRef]
  193. Rim, K.T.; Koo, K.H.; Park, J.S. Toxicological Evaluations of Rare Earths and Their Health Impacts to Workers: A Literature Review. Saf. Health Work 2013, 4, 12–26. [Google Scholar] [CrossRef]
  194. Yang, N.; Yang, J.; Liu, Y.; Fan, H.; Ji, L.; Wu, T.; Jia, D.; Ye, Q.; Wu, G. Impaired Learning and Memory in Mice Induced by Nano Neodymium Oxide and Possible Mechanisms. Environ. Toxicol. 2021, 36, 1514–1520. [Google Scholar] [CrossRef] [PubMed]
  195. Lehtinen, M.K.; Bonni, A. Modeling Oxidative Stress in the Central Nervous System. Curr. Mol. Med. 2006, 6, 871–881. [Google Scholar] [CrossRef]
  196. Peng, L.; Weiying, Z.; Xi, L.; Yi, L. Structural Basis for the Biological Effects of Pr(III) Ions: Alteration of Cell Membrane Permeability. Biol. Trace Elem. Res. 2007, 120, 141–147. [Google Scholar] [CrossRef] [PubMed]
  197. Mondal, D.; Rahman, M.M. Editorial: Exposure Pathways, Characterization and Risk Assessment of Chemical Contaminants in the Food Chain. Front. Environ. Sci. 2022, 10, 881334. [Google Scholar]
  198. Kurvet, I.; Juganson, K.; Vija, H.; Sihtmäe, M.; Blinova, I.; Syvertsen-Wiig, G.; Kahru, A. Toxicity of Nine (Doped) Rare Earth Metal Oxides and Respective Individual Metals to Aquatic Microorganisms Vibrio Fischeri and Tetrahymena Thermophila. Materials 2017, 10, 754. [Google Scholar] [CrossRef]
  199. Whitford, G.M. The Metabolism and Toxicity of Fluoride. Monogr. Oral Sci. 1996, 16, 151–153. [Google Scholar]
  200. Centers for Disease Control and Prevention CDC. Recommendations for Using Fluoride to Prevent and Control Dental Caries in the United States. MMWR Recomm. Rep. 2001, 50, 41–42. [Google Scholar]
  201. WHO (World Health Organization). Fluoride in Drinking Water. In Environmental Health Criteria Monograph; WHO: Geneva, Switzerland, 2004; Volume 227. [Google Scholar]
  202. Choi, A.L.; Sun, G.; Zhang, Y.; Grandjean, P. Fluoride, Developmental Fluoride Neurotoxicity. Environ. Health Perspect. 2012, 120, 1362–1369. [Google Scholar] [CrossRef]
  203. NRC (National Research Council). Fluoride in Drinking Water: A Scientific Review of EPA’s Standards; The National Academies Press: Washington, DC, USA, 2006. [Google Scholar]
  204. Barbier, O.; Arreola-Mendoza, L.; Del Razo, L.M. Molecular Mechanisms of Fluoride Toxicity. Chem. Biol. Interact. 2010, 188, 319–333. [Google Scholar] [CrossRef]
Figure 1. The relative germicidal effectiveness of UV radiation at different wavelengths (the so-called germicidal effectiveness curve—GEC) according to Gates [6] (red line) and DIN standard [35] (green line).
Figure 1. The relative germicidal effectiveness of UV radiation at different wavelengths (the so-called germicidal effectiveness curve—GEC) according to Gates [6] (red line) and DIN standard [35] (green line).
Nanomaterials 15 00562 g001
Figure 2. Illustration of 2-photon upconversion mechanisms with their respective quantum efficiencies (η) normalized to the incident power (1 W/cm2). Transitions between energy states are indicated by blue and red arrows, while green arrows indicate energy transfer. Adapted from [55], with permission from Elsevier; copyright 1990.
Figure 2. Illustration of 2-photon upconversion mechanisms with their respective quantum efficiencies (η) normalized to the incident power (1 W/cm2). Transitions between energy states are indicated by blue and red arrows, while green arrows indicate energy transfer. Adapted from [55], with permission from Elsevier; copyright 1990.
Nanomaterials 15 00562 g002
Figure 3. The general principle of antimicrobial photodynamic therapy (aPDT). A photosensitizer is attached to microorganisms (1) and, upon exposure to light of the suitable wavelength (2), becomes energized (3). The photosensitizer subsequently transfers energy to molecular oxygen, resulting in the generation of singlet oxygen and other reactive oxygen species capable of destroying microorganisms. Reprinted from [60] with permission of John Wiley and Sons; copyright 2025.
Figure 3. The general principle of antimicrobial photodynamic therapy (aPDT). A photosensitizer is attached to microorganisms (1) and, upon exposure to light of the suitable wavelength (2), becomes energized (3). The photosensitizer subsequently transfers energy to molecular oxygen, resulting in the generation of singlet oxygen and other reactive oxygen species capable of destroying microorganisms. Reprinted from [60] with permission of John Wiley and Sons; copyright 2025.
Nanomaterials 15 00562 g003
Figure 4. Illustration of the energy transfer sequence Nd3+ → Yb3+ → Tm3+ + Yb3+ → Gd3+ that enables 808 nm to UVC UC in NaGdF4:Yb,Tm@NaGdF4:Yb@NaGdF4:Yb,Nd@NaGdF4 core–multishell nanoparticles. Reprinted from [81] under a Creative Commons Attribution 4.0 international license (Creative Commons CC BY).
Figure 4. Illustration of the energy transfer sequence Nd3+ → Yb3+ → Tm3+ + Yb3+ → Gd3+ that enables 808 nm to UVC UC in NaGdF4:Yb,Tm@NaGdF4:Yb@NaGdF4:Yb,Nd@NaGdF4 core–multishell nanoparticles. Reprinted from [81] under a Creative Commons Attribution 4.0 international license (Creative Commons CC BY).
Nanomaterials 15 00562 g004
Figure 5. Energy level scheme of a free Pr3+ ion for (a) 4f2 configuration (note the vertical axis break) and (b) 4f5d configuration.
Figure 5. Energy level scheme of a free Pr3+ ion for (a) 4f2 configuration (note the vertical axis break) and (b) 4f5d configuration.
Nanomaterials 15 00562 g005
Figure 6. (a) Energy gap between 4f2 and 4f5d states in Pr3+ and (b) UV emissions from 4f5d to 4f2 transitions.
Figure 6. (a) Energy gap between 4f2 and 4f5d states in Pr3+ and (b) UV emissions from 4f5d to 4f2 transitions.
Nanomaterials 15 00562 g006
Figure 7. Blue-to-UV UC mechanisms of Pr3+ after first excitation with blue light populating the 3P2 level. (a) GSA/ESA process mediated by the 3P0 (blue arrows) and 1D2 (red arrows) levels. (b) GSA/APTE process mediated by the 3P0 (blue arrows) and 1D2 (red arrows) levels.
Figure 7. Blue-to-UV UC mechanisms of Pr3+ after first excitation with blue light populating the 3P2 level. (a) GSA/ESA process mediated by the 3P0 (blue arrows) and 1D2 (red arrows) levels. (b) GSA/APTE process mediated by the 3P0 (blue arrows) and 1D2 (red arrows) levels.
Nanomaterials 15 00562 g007
Figure 8. (a) Assessment of the biocidal effectiveness of polyvinyl alcohol/sodium alginate and Y2SiO5:Pr3+/polyvinyl alcohol/sodium alginate simulating S. aureus infection in vivo. (b) Illustration of the wound closure and (c) wound contraction rate on the 3rd, 5th, and 7th day. Reprinted from [152] with permission of the Royal Society of Chemistry; copyright 2023.
Figure 8. (a) Assessment of the biocidal effectiveness of polyvinyl alcohol/sodium alginate and Y2SiO5:Pr3+/polyvinyl alcohol/sodium alginate simulating S. aureus infection in vivo. (b) Illustration of the wound closure and (c) wound contraction rate on the 3rd, 5th, and 7th day. Reprinted from [152] with permission of the Royal Society of Chemistry; copyright 2023.
Nanomaterials 15 00562 g008
Figure 9. (a) An illustration of the proof-of-concept biofilm deactivation with Pr3+-doped Y2Si2O7 and Pr3+,Yb3+,Tm3+-doped powders. (b) Viability of biofilm (S. aureus; A. baumannii, and C. albicans) cells after irradiation with 447 nm laser light (800 mW/cm2) in the presence of un-doped Y2Si2O7; Y2Si2O7:Pr3+, and Y2Si2O7:Pr3+,Tm3+,Yb3+ and without them (control, 100% represents viability of non-irradiated cells). (c) Effectiveness of generating ROS in biofilm irradiated by a 447 nm laser diode (800 mW/cm2) in the presence of un-doped Y2Si2O7; Y2Si2O7:Pr3+; and Y2Si2O7:Pr3+,Tm3+,Yb3+ phosphors. Reprinted from [157] under a Creative Commons Attribution 3.0 international license (Creative Commons CC BY).
Figure 9. (a) An illustration of the proof-of-concept biofilm deactivation with Pr3+-doped Y2Si2O7 and Pr3+,Yb3+,Tm3+-doped powders. (b) Viability of biofilm (S. aureus; A. baumannii, and C. albicans) cells after irradiation with 447 nm laser light (800 mW/cm2) in the presence of un-doped Y2Si2O7; Y2Si2O7:Pr3+, and Y2Si2O7:Pr3+,Tm3+,Yb3+ and without them (control, 100% represents viability of non-irradiated cells). (c) Effectiveness of generating ROS in biofilm irradiated by a 447 nm laser diode (800 mW/cm2) in the presence of un-doped Y2Si2O7; Y2Si2O7:Pr3+; and Y2Si2O7:Pr3+,Tm3+,Yb3+ phosphors. Reprinted from [157] under a Creative Commons Attribution 3.0 international license (Creative Commons CC BY).
Nanomaterials 15 00562 g009
Figure 10. The relative germicidal efficiency curve (green-shaded spectral region; according to DIN [35]) and 4f5d UC emission spectra overlap in certain Pr3+-activated hosts often used in Pr3+-based blue-to-UVC UC.
Figure 10. The relative germicidal efficiency curve (green-shaded spectral region; according to DIN [35]) and 4f5d UC emission spectra overlap in certain Pr3+-activated hosts often used in Pr3+-based blue-to-UVC UC.
Nanomaterials 15 00562 g010
Table 1. Some examples of antimicrobial photodynamic therapy (aPDT) applications, including the materials used, the UC hosts, the dopants that play a role in UC, the absorption and emission wavelengths, the germicidal effect that was achieved, and a literature reference.
Table 1. Some examples of antimicrobial photodynamic therapy (aPDT) applications, including the materials used, the UC hosts, the dopants that play a role in UC, the absorption and emission wavelengths, the germicidal effect that was achieved, and a literature reference.
MaterialUC Host
Doping Ions
(Concentrations)
Excitation and EmissionAntimicrobial EffectRef.
CuS-decorated NaYF4 nanoparticles coated with methylene blue doped silica and grafted with chitosanNaYF4
Mn2+/Yb3+/Er3+ (30/18/2 mol%)
Exc. 980 nm
emission in red (651 nm)
Synergistic photothermal and photodynamic therapy
effective against Gram-positive S. aureus and Gram-negative E. coli.
[69]
NaYF4@mSiO2
mSiO2 (mesoporous silica) shell loaded with hydrophobic photosensitizer SiPc (silicon 2,9,16,23-tetra-tert-butyl-29H,31H-phthalocyanine dihydroxide)
Cubic NaYF4
Yb3+/Er3+
(20/2 mol%)
Exc. 976 (power density of 2 W/cm2)
emission in green (520–560 nm) and red (640–680 nm)
Complete eradication of E. coli and seven-order-of-magnitude decrease in colony-forming units of S. aureus.[70]
Roussin’s black salt (RBS)-loaded UCNPs
NaGdF4@mSiO2@qC
(qC—quaternized
ammonium chitosan)
NaGdF4
Yb3+/Tm3+
(25/0.3 mol%)
Exc. 980 nm (1 W)
emission in UV (290, 345, and 362 nm), blue (450 nm and 474 nm), red (574 nm and 643 nm), and near-infrared (807 nm)
Nitric oxide triggered antibacterial activity against methicillin-resistant S. aureus (MRSA) and E. coli in vitro and in vivo.[71]
N-octyl chitosan-coated NaYF4:Yb,Er@NaYF4 core–shell nanoparticles loaded with the zinc phthalocyanine photosensitizerhexagonal
NaYF4
Yb3+/Er3+
(concentrations are not provided)
Exc. 980 nm emission in green (520–560 nm) and red (640–680 nm)Effective against methicillin-resistant S. aureus (MRSA) and E. coli. Effective treatment of the MRSA-infected abscesses in deep tissue (1 cm).[72]
Rose Bengal (photosensitizer)-loaded LiYF4 capped with polyvinylpyrrolidoneLiYF4
Yb3+/Er3+
(concentrations are not provided)
Exc. 980 nm (power density of 1 W/cm2) Emission in green (520–560 nm) and red (640–680 nm)Effective in deep tissue infections; used with methylene blue for aPDT. Decline of 4.72 log10 in viability of drug-resistant Acinetobacter baumannii at a dose of 50 μg mL−1 UCNPs-PVP-RB.[73]
Rose Bengal (photosensitizer)-loaded NaYF4:Yb,Er@NaGdF4:Nd@SiO2 core–shell nanoparticlesNaYF4
Yb3+/Er3+
(18/2 mol%)
Exc. 980 nm (power density of 1 W/cm2); emission in green (520–560 nm) and red (640–680) nmEffective against methicillin-resistant S. aureus (MRSA) and E. coli.[74]
D-TiO2/Au@SiO2@Y2O3:Yb3+,Er3+ with an antibiotic drug Ampicillin sodium covalently linked to the nanoparticles by a (3-
glycidyloxypropyl)trimethoxysilane monolayer linker
Y2O3
Yb3+/Er3+
(concentrations are not provided)
Exc. 980 nm (power density of 0.68 W/cm2); emission in green (520–560 nm) and red (640–680 nm)Effective against methicillin-resistant S. aureus (MRSA) and E. coli.[75]
NaErF4:Tm3+@NaYF4:Yb3+-Chlorin e6-Mn(CO)5Br@SilaneNaErF4:Tm3+
Er3+/Tm3+
(concentrations are not provided)
NaYF4
Yb3+
(concentration is not provided)
Exc. 980 nm (power density of 1 W/cm2). Emission in red (660 nm)At 150 μg/mL, the therapy results in inhibition of over 70% of E. coli and S. aureus, while at 200 μg/mL, it inhibits approximately 90% of both bacteria strains. Additionally, an anti-inflammatory effect is observed.[65]
Heavy metal-free organic photosensitizer attached to the NaGdF4:Nd3+/Tm3+/Yb3+@NaGdF4 core–shell nanoparticles coated with a phospholipid bilayerNaGdF4
Yb3+/Nd3+/Tm3+
(25/1/0.5 mol%)
Exc. 808 nm (power density of 140 mW/cm2 or 3.2 W/cm2). Emission in UV (340, and 360 nm) and blue (450 nm and 480 nm).HeLa cells are efficiently destroyed via 808 nm laser irradiance of 140 mW/cm2 for 3 min (<30% cell viability) or via 3.2 W/cm2 for 6 min (<10% cell viability).[76]
(CTAB-coated NaYF4:Yb/Tm)@ZnOβ-NaYF4
Yb3+/Tm3+
(18/5 mol%)
Exc. LED 970 nm (power of
12 mW/cm2). Emission in UVA (345 and 362 nm) and blue (451 nm and 475 nm)
CFU reduction of S. aureus WCH-SK2-SCV of 82.6% and S. aureus WCH-SK2 of 78.8% is demonstrated.[77]
Table 2. Pr3+-activated materials showing 4f5d emission in UVC. A more comprehensive list is given in Refs. [90,91].
Table 2. Pr3+-activated materials showing 4f5d emission in UVC. A more comprehensive list is given in Refs. [90,91].
HostCrystal StructureSpace GroupMain Emission Peaks (nm)Reference Cutoff   Phonon   Energy   ω m a x (cm−1) with [Reference]
Cs2NaYCl6CubicFm3m263, 277, 301, 314[92]284 [93]
LaOITetragonalP4nmm300[94]<430 [95] *
KCaF3HexagonalPnma257[96]412 [97]
RbCaF3CubicPm-3m261[96]486 [98]
CsCaF3CubicPm-3m250, 273[96]449 [98]
LiLuF4TetragonalI41/amd223–281[99]445 [100]
CaSO4OrthorhombicAmma223, 234, 250, 255[101]1185 [102]
Cs2NaYF6CubicFm-3m250, 270[103]467 [93]
YBO3HexagonalP63/m263, 275[89]1368 [104]
La2CaB10O19MonoclinicC2279, 334 [105]1493 [106]
LuPO4TetragonalI41/amd235, 246, 263, 274[87]1161 [107]
YPO4TetragonalI41/amd232, 244.5, 261.6, 271[108]1149 [107]
NaCaPO4OrthorhombicPn21a251, 261, 282[109]1080 [110]
K3Lu(PO4)2TrigonalP-3253, 282, 315[87,111]1147 [112] **
Sr3(PO4)2TrigonalR3-m231, 269[113]1072 [114]
Sr3Y(PO4)3CubicI-43d248, 278[115]1080 [116]
Ba3Y(PO4)3CubicI-43d250, 280[115]1044 [117]
Ca9Y(PO4)7TrigonalR-3c240, 275[118]1125 [118]
LiLuSiO4OrthorhombicPnma268, 283, 316[87]980 [119] ***
Li2SrSiO4HexagonalP3121265, 315[120]884 [121]
LiY9(SiO4)6O2HexagonalP63/m268[122]958 [123]
Lu2SiO5MonoclinicC2/c275, 313[124]970 [125]
X2-Y2SiO5MonoclinicC2/c270, 282, 308[126]971 [127]
* LaOI has a lower phonon energy than LaOBr due to the larger atomic mass of I than Br; ** phonon energy of Rb3Lu(PO4)2; *** phonon energy of LiYbSiO4.
Table 3. Literature data on bactericidal effects induced by Pr3+ UC UVC radiation.
Table 3. Literature data on bactericidal effects induced by Pr3+ UC UVC radiation.
UC MaterialEmissionExcitationAntimicrobial EffectRef.
β-NaYF4:Pr3+/Li+ and β-NaYF4: Pr3+/Li+@BiOCl composite
(dopant concentrations are not provided)
UVC (253 nm, 259 nm, 284 nm)444 nm Antimicrobial effect of β-NaYF4:Pr3+/Li+ under 444 nm excitation demonstrated. With the β-NaYF4: Pr3+/Li+@BiOCl composite, the effect is significantly improved (visible light excitation ≥ 420 nm kills 99.99% of E. coli in 180 min—aPDT effect).[153]
Lu7O6F9:Pr3+
(1 mol%)
UVC 260 nm447 nmInactivation of E. coli implicated by the authors.[154]
NaYF4:Pr3+/Yb3+
(2/10 mol%)
LiYF4:Pr3+/Yb3+
(1/10 mol%)
UVC 275 nm447 nmSignificant denaturation of a double strand DNA after exposure to 447 nm radiation for 20 to 40 min. Material’s additional functionality is luminescence imaging in the NIR-II spectral region.[155]
Cs2NaYF6:Pr3+
(1 mol%)
UVC afterglow at 250 nmX-rayFollowing 16 min of X-ray irradiation, the sample is placed near to a plate containing a colony of the Gram-negative bacteria Pseudomonas aeruginosa. The viability of around 40% of bacteria is seen under the UVC afterglow of this material.[103]
Y2SiO5:Pr3+
(1 mol%)
Y2SiO5:Pr3+/Gd3+
(1/1 mol%)
Y2SiO5:Pr3+/Gd3+/Li+
(1.2/1.2/7.2 mol%)
Y2SiO5:Pr3+/Li+
(1.2/7.2 mol%)
UVC 280 nm, additional emission at 318 nm with samples containing Gd3+“daylight” fluorescent lightingInactivation of B. subtilis spores on dry phosphor-coated surfaces (best results with Pr3+/Gd3+/Li+-doped material). Inhibition of P. aeruginosa biofilms grown on the coated surfaces.[142]
Ca2SiO4:Pr3+
(dopant concentration is not provided)
UVC 247 nm450 nm laserInactivation of B. subtilis.[156]
Y2SiO5:Pr3+ composite film with polyvinyl alcohol (PVA) and sodium alginate (SA)UVC 280 nm455 nm and
white LED
Inhibition of Gram-positive S. aureus and Gram-negative E. coli Pseudomonas aeruginosa bacteria.[152]
Y2Si2O7:Pr3+/Tm3+/Yb3+
(1.2/0.5/5 mol%)
UVC (278 nm) + UVB (308 nm) + UVA (370 nm)447 nm laser, 800 mW/cm2The viability of planktonic cultures of A. baumannii, S. aureus, and C. albicans.[157]
Li2SrSiO4:Pr3+
(dopant concentration is not provided)
Two broad peaks (265 nm and 320 nm). 450 nm laser, 1 WInactivation of Bacillus subtilis: after 300 s of irradiation, the mortality rate reached 90%, and after 600 s of irradiation, almost all bacteria died.[120]
Li2CaGeO4:Pr3+
(1 mol%)
UVC+UVB (~240–330 nm)450 nm laser, 1 WComplete inactivation of S. aureus bacteria in 30 min.[158]
Li2SrGeO4
(1 mol%)
UVC+UVB (~240–330 nm)450 nm laser diode, 0.6 WInactivation of Staphylococcus aureus, Salmonella enterica, Klebsiella pneumoniae, and Escherichia coli in 40 to 80 min.[159]
Table 4. SOC values for Pr3+ materials whose emissions are shown in Figure 10; calculated using Equation (3).
Table 4. SOC values for Pr3+ materials whose emissions are shown in Figure 10; calculated using Equation (3).
LaPO4YPO4Lu7O6F9YBO3X2-Y2SiO5Sr3(BO3)2Lu2SiO5β-Y2Si2O7
SOC0.2820.3570.6740.5290.5330.5350.5040.477
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

Dramićanin, M.D.; Brik, M.G.; Antić, Ž.; Bănică, R.; Mosoarca, C.; Dramićanin, T.; Ristić, Z.; Dima, G.D.; Förster, T.; Suta, M. Pr3+ Visible to Ultraviolet Upconversion for Antimicrobial Applications. Nanomaterials 2025, 15, 562. https://doi.org/10.3390/nano15070562

AMA Style

Dramićanin MD, Brik MG, Antić Ž, Bănică R, Mosoarca C, Dramićanin T, Ristić Z, Dima GD, Förster T, Suta M. Pr3+ Visible to Ultraviolet Upconversion for Antimicrobial Applications. Nanomaterials. 2025; 15(7):562. https://doi.org/10.3390/nano15070562

Chicago/Turabian Style

Dramićanin, Miroslav D., Mikhail G. Brik, Željka Antić, Radu Bănică, Cristina Mosoarca, Tatjana Dramićanin, Zoran Ristić, George Daniel Dima, Tom Förster, and Markus Suta. 2025. "Pr3+ Visible to Ultraviolet Upconversion for Antimicrobial Applications" Nanomaterials 15, no. 7: 562. https://doi.org/10.3390/nano15070562

APA Style

Dramićanin, M. D., Brik, M. G., Antić, Ž., Bănică, R., Mosoarca, C., Dramićanin, T., Ristić, Z., Dima, G. D., Förster, T., & Suta, M. (2025). Pr3+ Visible to Ultraviolet Upconversion for Antimicrobial Applications. Nanomaterials, 15(7), 562. https://doi.org/10.3390/nano15070562

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