Next Article in Journal
Natural Characteristics Analysis of a Dual-Rotor System with Nonparametric Uncertainty
Next Article in Special Issue
Commissioning of Bunch Compressor to Compress Space Charge-Dominated Electron Beams for THz Applications
Previous Article in Journal
Micro-Vibration Signal Denoising Algorithm of Spectral Morphology Fitting Based on Variational Mode Decomposition
Previous Article in Special Issue
Acceleration of an Electron Bunch with a Non–Gaussian Transverse Profile in Proton-Driven Plasma Wakefield
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Highly Enriched Uranium-Free Medical Radioisotope Production Methods: An Integrative Review

by
Bruno Silveira Nunes
1,2,*,
Enio Rodrigo Fernandes Rodrigues
2,
Jonathan Alexander Prestes Fruscalso
2,
Roger Pizzato Nunes
2,3,
Alexandre Bonatto
1,2 and
Mirko Salomón Alva-Sánchez
1,2
1
Graduate Program in Information Technology and Healthcare Management, Federal University of Health Sciences of Porto Alegre (UFCSPA), Porto Alegre 90050-170, RS, Brazil
2
Beam Physics Group, Federal University of Health Sciences of Porto Alegre (UFCSPA), Porto Alegre 90050-170, RS, Brazil
3
Electrical Engineering Department, Federal University of Rio Grande do Sul (UFRGS), Porto Alegre 90035-190, RS, Brazil
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(24), 12569; https://doi.org/10.3390/app122412569
Submission received: 15 October 2022 / Revised: 12 November 2022 / Accepted: 6 December 2022 / Published: 8 December 2022
(This article belongs to the Special Issue Advanced Technologies of Particle Accelerators and Their Applications)

Abstract

:
The ever-growing need for radiopharmaceuticals, i.e., compounds containing pharmaceutical drugs and radioisotopes used for medical diagnostic imaging (SPECT/PET scan) and treating neoplasms, is significantly leading to an increased demand for such substances in hospitals and clinics worldwide. Currently, most large-scale productions of radioisotopes required for radiopharmaceuticals are carried out in research reactors, via the fission of highly enriched uranium. However, because large amounts of radioactive waste are produced as byproducts in this process, new greener methods are needed for radioisotope production. This work presents an integrative literature review and summarizes enriched uranium-free methods for radioisotope production, accomplished through the adoption of new reaction routes, distinct acceleration technologies, or by using other physical processes. This review considered forty-eight studies published from 2010 to 2021 on three established virtual databases. Among these selected works, a cyclotron is the most adopted HEU-free method for radioisotope production, and 44 Sc, 68 Ga, and 99 m Tc are the medical radioisotopes most often reported as produced by using the investigated HEU-free production methods.

1. Introduction

Hospitals, clinics, and health centers make daily use of a wide range of highly complex diagnoses and treatments resulting from technological advances achieved in the last decades, with nuclear medicine being an area that benefits from such advances. In nuclear medicine, radiopharmaceuticals, which are compounds containing radionuclides (or radioisotopes) bound to pharmaceutical drugs [1], are administered to patients via intravenous or oral injections. Depending on their properties, these organic molecules allow the radionuclides to be fixed in specific human body organs. The radioisotope corpuscular and/or the electromagnetic radiation it emits in this organ is used for imaging or treatment purposes [2]. Radionuclides emitting either β + particles and/or γ -rays are usually employed in diagnoses. On the other hand, Auger electrons and β -emitter radioisotopes are administered in therapy [3]. Medical radioisotopes must have half-lives long enough to allow image acquisition or therapeutic effects, but their half-lives must be short enough to minimize the patient’s exposure to ionizing radiation [4]. In this sense, not all radioisotopes are appropriate to be used in radiopharmaceuticals.
Some examples of medical radioisotopes are the 99 m Tc (half-life of approximately 6 h), 18 F (half-life of 1.83 h), and 64 Cu (half-life of 12.7 h). Moreover, 99 m Tc, the most used radioisotope in nuclear medicine, is a γ -ray-emitter radioisotope; its emitted radiation is detected through a gamma camera in a SPECT (single photon emission tomography) equipment, enabling internal imaging of the patient for diagnostic purposes [5]. Routine clinical applications of 99 m Tc include bone scintigraphy [6], myocardial perfusion [7], ventriculography [8], brain [9], sentinel node [10], immunoscintigraphy [11], blood-pool labeling [12], sulfur colloid [13], and Meckel’s diverticulum [14] scan procedures. On the other hand, 18 F is a β + -emitter radioisotope applied in positron emission tomography (PET) for the early detection and diagnosis of Hodgkin lymphoma [15], and colorectal, head and neck, lung, cervical, and ovarian cancers [16]. Finally, 64 Cu is a β (39 %) and β + (61 %) emitter used for both PET imaging [17,18,19] and cancer therapy [20].
Currently, most of the radioisotope demand is met by two production methods: the fission of highly enriched uranium (HEU) in research reactors, or the acceleration of electrically charged particles (usually protons) in cyclotrons, toward gaseous, liquid, or solid targets, depending on the radioisotope to be produced [21]. Cyclotrons are most often utilized to produce short-life, neutron-deficient, β + -emitter radioisotopes, which are mainly used for PET scans. Because of their short half-lives, these radioisotopes need to be used right after production. Hence, they are mostly produced in medical cyclotrons and are immediately utilized in the same sites in which these facilities are installed [22]. On the other hand, long-lived β and γ -ray-emitter radioisotopes, generally applied to internal radiotherapy and SPECT scans, respectively, are primarily produced in research reactors. In addition to enabling large-scale production, research reactors also offer logistic advantages. Because most of the radioisotopes produced in these facilities have longer half-lives, if compared to those typically produced in cyclotrons [22], they can be widely distributed before being used. However, as it will be soon discussed, this comes at the cost of transporting highly radioactive material over long distances.
In a research reactor, a flux of up to 10 14 neutrons/ c m 2   s hits an HEU target [23], starting the uranium fission. This process produces over forty distinct radioisotopes [24], including the 99 Mo (half-life of 66 h), which later decays to 99 m Tc in a 99 Mo/ 99 m Tc generator. Due to the molybdenum’s large half-life, this generator can be transported all over the world to support the worldwide medical demand of 99 m Tc. Despite the significant advantages of the radioisotope production scale, multiple substantial concerns are associated with the fission of HEU in research reactors. The first (and most dangerous) one is the generation of large amounts of environmentally hazardous radioactive waste, which can take from decades to hundreds of years to decay. Dealing with radioactive waste often requires relocating and storing it in a deep–underground repository [25], which leads to storage space/safety problems since the waste can be active for hundreds of years. Another issue caused by having most of the radioisotope production centralized in research reactors is the consequent need of transporting highly radioactive material over long distances [26]. Although this activity has been safely and efficiently conducted for several decades [27], there are issues associated with it, such as the higher production costs and logistics complexities, which are even more critical for developing countries [28]. Finally, as mentioned by (Van Noorden, R. 2013) [29], the shutdown of two research reactors in 2009, the Chalk River reactor (in Canada) and the Petten high flux nuclear reactor (in the Netherlands), made it clear that the world’s radioisotope supply chain was fragile, as it greatly relied on a few research reactors built in the 1950s and 1960s. In those times, these two reactors were responsible for producing most of the world’s supply of 99 m T; thus, the shutdowns created a shortage of this radioisotope. This crisis motivated the research into 99 Mo production in cyclotrons via proton irradiation, and in linear accelerators via the bremsstrahlung γ -ray reaction route 100 Mo( γ ,n) 99 Mo [29].
In addition to the aforementioned issues associated with using research reactors, international nuclear non-proliferation policies emphasize the need of eliminating the use of HEU for medical radioisotope production [30]. Hence, developing alternatives to the HEU-based production methods, i.e., HEU-free production routes through different acceleration technologies and/or physical processes, is a crucial step toward finding a greener solution to meet the ever-growing demand for medical radioisotopes. In order to contribute to this relevant topic, this integrative review [31] aims to identify and synthesize the recently reported HEU-free medical-radioisotope production methods, as well as the medical radioisotopes produced by using such methods.
This review is organized as follows. Section 2 presents an overview of the method adopted for the literature review and describes the process used to analyze the results. In Section 3, the results obtained are presented and discussed. Finally, in Section 4, the conclusions and future directions are presented.

2. Materials and Methods

This study aims to identify which radioisotopes used in medicine have recently received interest from the scientific community and the methodologies used in their production. Thus, this research has two guiding questions: “What alternative, HEU-free methods have been recently used to produce medical radioisotopes?” and “What medical radioisotopes can be produced by these alternative technologies?”.
The literature review was performed in January 2022 on the databases Virtual Health Library (VHL) [32], Science Direct [33], and Scopus [34], with the following search strings: “radiopharmaceutical production” OR “radioisotope production”. The search filters used were the following: peer-reviewed papers; English-written; available electronically; and published from 2010 to 2021. By using these filters, 91 papers were selected on VHL, 604 on Science Direct, and 294 on Scopus, resulting in 989 papers. After removing duplicate papers, this number was reduced to 927 unique papers. After a preliminary analysis, it was observed that, rather than reporting their own experimental radioisotope production, many of these publications only cited the productions from other works (eventually comparing them to their own simulation results) or estimated radioisotope reaction cross-sections, without informing the production itself. Since filtering this situation is non-trivial, once it involves context awareness, the 927 selected papers were manually analyzed. From this analysis, papers that did not report their own experimental radioisotope production, or did not present explicit radioisotope activity—in units of Bq or Ci—thick target yields, or saturation yields (which could be used to calculate the activity), were excluded. Using this methodology, from the 927 papers, 48 were selected [35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82], and their data are presented in Table A1, Table A2, Table A3 and Table A4, in chronological order. In particular, in Table A2, Table A3 and Table A4, experimental results are reported with the same precision (number of decimal digits) informed by their respective authors. Moreover, quantities calculated from these results were rounded according to the number of significant digits.
With regard to the amount of radioisotope produced, all data collected from these papers were converted to activities at the end of the bombardment in order to present all results in the same physical unit. Each selected article presented the information differently, because the production technology used was not the same, or because different types of measurements were conducted.
Most of the selected studies utilizing cyclotron accelerators presented their production results in one of three ways: thick target yield, saturation yield, or activity at the end of the bombardment. In order to compare the production with studies that utilized other accelerator technologies or physical processes, all measurements or calculated results presented in these papers were converted to activities at the end of the bombardment, by using Equations (1) and (2) [83]:
A EOB = TTY i λ ( 1 e λ t ) ,
A EOB = SY i ( 1 e λ t ) ,
where A EOB is the activity at the end of the bombardment [ M Bq ], t is the irradiation time [ h ], TTY is the thick target yield [ M Bq / A h ], λ is the decay constant of the produced radioisotope [ h 1 ], i is the beam current [ A ], and SY is the saturation yield [ M Bq / A ].
Some authors performed experimental activity measurements hours after the end of the bombardment. In those cases, the A EOB was estimated with Equation (3) using the measurement activity A T , the time after irradiation T [ h ], and the decay constant λ .
A EOB = A T e λ T .

3. Results and Discussion

Table A1 and Figure 1a show that 81.3% of the selected papers describe the use of cyclotrons for radioisotope production. The reason behind this large number is that cyclotron is a well-established technology. The use of hospital cyclotrons to produce radioisotopes began in the 1950s [84], over 70 years ago, and the largest medical centers in the world already have cyclotrons. Globally, there are more than 1500 cyclotrons [85]. Protons are usually the types of particles accelerated in cyclotrons, justifying their large appearance numbers (34), as indicated in Table A1 and Figure 1b. In addition to cyclotrons, research reactors (without the presence of enriched uranium), linear particle accelerators (LINACS), plasma focus devices, and tandetrons have also been used for radioisotope production in the selected papers. These four production technologies add up to only 18.7% of the papers, as shown in Figure 1a.
Figure 2 shows that 53 unique radioisotopes were produced in the selected papers; the most cited medical radioisotopes were 44 Sc, 66 Ga, 99 m Tc, 68 Ga, 99 Mo, and 89 Zr. Together, they appear in 25% of the selected papers.
As shown in Figure 2, the most cited (7 distant citations) radioisotope is the β + -emitter radioisotope 44 Sc. Its half-life is often described to be 3.97 h, but a recent study reported a half-life of 4.04 h [86]. This radioisotope was investigated for PET imaging [87] and might be a suitable alternative to 68 Ga (with a half-life of 1.13 h) due to its half-life, almost four times greater than 68 Ga, which allows the synthesis of a variety of radiotracers with longer pharmacokinetic profiles [88]. Additionally, 44 Sc has been proposed as a matching pair of 47 Sc (with a half-life of 80.4 h, a β -emitter that has been proposed for radioimmunotherapy [89]) for PET imaging, dosimetry estimation, and the assessment of radionuclide therapeutic responses [89,90,91]. Currently, there are two main routes for 44 Sc production, i.e., via 44 Ti/ 44 Sc generators, but due to difficulties in the production of the parent isotope ( 44 Ti), in the chemical separation, and the purification process, the applicability of this generator is very limited, making its clinical implementation difficult [92]. Cyclotron production is the other main source of the 44 Sc radioisotope. In the selected papers, the production occurred via irradiation on 44 Ca, 42 Ca, or nat Ca targets through 44 Ca(p,n) 44 Sc, 44 Ca(d,2n) 44 Sc, 42 Ca( α , np+pn) 44 Sc, and nat Ca(p,n) 44 Sc reaction routes, as indicated in Table A1.
With three distinct citations, 99 Mo is a β -emitter radioisotope without direct application in medicine. However, 99 Mo decays into 99 m Tc, the most used radioisotope for SPECT imaging, as mentioned in Section 1. It makes 99 Mo and 99 m Tc the most relevant pair of radioisotopes in nuclear medicine, justifying a total of seven distinct paper citations (four papers cited 99 m Tc and three cited 99 Mo).
Figure 2 also shows that two radioisotopes of gallium, namely 68 Ga and 66 Ga, had seven citations combined. Moreover, 66 Ga (with a half-life of 9.49 h) is a β + -emitter used in PET. Its long half-life makes it possible to perform next-day PET imaging, improving the image contrast [93]. 66 Ga has been tested for studies of slow dynamic processes, such as lymphatic transport [94], and imaging of tumor angiogenesis using monoclonal antibodies [95]. Moreover, there have been reports of the use of 66 Ga in a 66 Ga-labeled somatostatin as an imaging agent for receptor-positive tumors [96] and in preclinical imaging of the GRPR expression in prostate cancer [97]. Table A1 shows that, in the papers found in our review, 66 Ga was produced in cyclotrons via irradiation on zinc targets through the nat Zn(p,x) 66 Ga and 66 Zn(p,n) 66 Ga reaction routes. Moreover, as used in positron emission tomography, 68 Ga is a β + -emitter radioisotope applied in the diagnostics of prostate cancer [98] and neuroendocrine neoplasms [99]. It is also produced in cyclotrons via irradiation on zinc targets through the nat Zn(p,x) 68 Ga and 68 Zn(p,n) 68 Ga reaction routes, as shown in Table A1. Another possibility to produce this radioisotope is by a 68 Ge/ 68 Ga generator [100,101], which can consistently supply 68 Ga for more than a year due to the 68 Ge half-life of 271 days.
The last radioisotope with three citations in distinct papers is 89 Zr (half-life of 78.4 h), a β + -emitter, which is an effective imaging tool for antibody or immune-based PET, referred to as “immuno-PET” [102], which represents by far the widest field of the 89 Zr medical application. However, there are studies on 89 Zr-labeled nanoparticles (NPs) with promising results in tumor detection, drug monitoring, inflammation imaging, tumor-associated macrophages and sentinel lymph mapping [103]. Table A2 shows that the three papers produced 89 Zr in a cyclotron via irradiation on yttrium targets through the 89 Y(p,n) 89 Zr reaction route.
As shown in Table A2, Table A3 and Table A4, there is a wide range of A EOB values, spanning multiple orders of magnitudes (from a few Bq to 10 14   Bq ). This extreme variation is a consequence of the distinct production routes, technologies, irradiation times, and electric currents or neutron fluxes used for producing each radioisotope reported in these tables. As indicated in Table A2 n° 32, for the 44 Sc production, the activity obtained with a natural target is two orders of magnitude lower than the one obtained with an enriched target. If compared to an equivalent enriched target, a natural target will contain only a fraction of the isotopes available for a given radioisotope production route. Hence, a much lower yield is expected for the natural material, if compared to the enriched target.

4. Conclusions and Future Directions

An integrative review of alternative methods for radioisotope production was carried out. Papers containing experimental production results, published from 2010 to 2021, were considered. A total of 48 papers were selected and analyzed. Among these works, cyclotrons were the most adopted particle accelerator technologies for radioisotope production, followed by HEU-free research reactors. Combined, these two methods were adopted in nearly 94% of the papers evaluated in this review.
Regarding the radioisotopes, the most-cited one was 44 Sc, with seven distinct citations, followed by 68 Ga and 99 m Tc with four citations, and 66 Ga, 99 Mo, and 89 Zr with three citations each. All of these radioisotopes are applied in PET ( 44 Sc, 68 Ga, 66 Ga, 89 Zr), or SPECT ( 99 Mo, 99 m Tc) diagnoses.
It is clear that new compact, environmentally safe, and cost-effective technologies for producing radioisotopes are needed, and that there is interest from the scientific community in developing new, HEU-free production routes that mitigate the generation of radioactive waste. Moreover, new, decentralized production methods could minimize possible future problems similar to the shortage crisis of 99 Mo/ 99 m Tc generators in 2009. As shown in this review, HEU-free medical-radioisotope production methods have been adopted and reported in the last decade. However, more research in this area is crucial for attaining greener and high-yield radioisotope production methods, capable of competing with the current uranium-based standard production techniques.
A possible alternative would be producing radioisotopes via reaction routes triggered by bremsstrahlung γ -rays, obtained, for example, from the deceleration of high-energy electron beams. Depending on the target material, it is possible to estimate the energy range required for the photons to trigger photonuclear reactions capable of producing the radioisotope of interest. Moreover, as in this process, the radioisotope of interest is a product of the reaction, rather than a byproduct, as in the case of nuclear fission, the adoption of this process would mitigate the radioactive waste. Despite being reported in the literature as a possible scheme for producing radioisotopes [104,105], such as, for example, the 99 Mo (which decays to 99 m Tc) [105,106], no experimental works were found among the papers selected for this review according to the aforementioned criteria.
The high-energy electron beams required to generate bremsstrahlung photons capable of producing radioisotopes via photonuclear reactions might be soon attainable by using laser wakefield accelerators (LWFAs) [107,108]. In this scheme, under the proper conditions, an intense laser pulse propagating in plasma drives a high-amplitude wakefield on its trail, which can be used as a compact accelerating structure for the plasma electrons (self-injection), or for an externally injected electron beam. LWFAs are capable of producing eight GeV electron beams at the exit of a 20 cm-long plasma capillary [109]. Moreover, stable long-term kHz operations of LWFAs have recently been reported [110]. In addition to the low amount of radioactive waste produced via bremsstrahlung γ -rays reaction routes, the compactness of this technology could allow for local radioisotope production, mitigating the aforementioned transportation problems associated with other technologies that require centralized production.
Despite the LWFA being a promising technology [111], recent studies [106,112,113] indicate that, even if high-energy, high-repetition-rate laser systems are adopted, due to the low yield provided by the photon activation process, very long irradiation times would still be required to produce a given radioisotope in quantities that could supply typical daily doses for medical applications. Hence, laser technology development, as well as further studies on LWFA optimization for radioisotope production—by using, for example, artificial intelligence algorithms [114,115]—may enable the development of greener, non-centralized radioisotope production processes.
Another possible laser-based radioisotope production method is the target normal sheath acceleration (TNSA) [112,116,117], in which an ultra-intense–ultra-short laser pulse interacts with a solid target. A fraction of the laser’s electromagnetic field is transferred to the target, creating plasma and heating electrons to relativistic temperatures. The electrons expand and exit the target, forming a charge separation localized next to its surface. This process generates a strong electrostatic field, perpendicular to the target surface, capable of accelerating ions from the target to several MeVs [118,119]. This ion irradiation can be used to produce radioisotopes via (p,x) and (d,x) reaction routes. The first experimental TNSA production of 99 m Tc through the 100 Mo(p,2n) 99 m Tc reaction route occurred in 2013 [120]. In the experiment, a 1 mm thick aluminum target was illuminated by a 50 J, 1 ps, 5 × 10 20 W/cm 2 laser pulse. For a single shot, a production of 8.25 kBq of 99 m Tc was estimated. Moreover, simulation results with the 3D PIC simulation code Mandor [121] show that a 10 J, 100 fs laser might produce 1.93 kBq of 99 m Tc in a single shot through the same reaction route, 100 Mo(p,2n) 99 m Tc. Hence, in principle, it should be possible to produce 300 GBq of 99 m Tc after 6 h of continuous 10 kHz laser irradiation on a highly enriched 100 Mo target [122].
Despite the encouraging progress in both experimental and theoretical research studies on laser-based radioisotope production, there are still many improvements to be made before laser-driven radioisotope production facilities become viable for medical applications.

Author Contributions

Conceptualization, R.P.N., A.B. and M.S.A.-S.; methodology, B.S.N., E.R.F.R. and J.A.P.F.; writing—original draft preparation, B.S.N.; writing—reviewing and editing, B.S.N., A.B., M.S.A.-S. and R.P.N.; supervision, R.P.N., A.B. and M.S.A.-S.; project administration, M.S.A.-S. All authors have read and agreed with the published version of the manuscript.

Funding

This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior—Brasil (CAPES) (grant 88887.620985/2021-00), and by the Fundação de Amparo à Pesquisa do Estado do Rio Grande do Sul (FAPERGS) (grant 21/2551-0002027-0).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Sílvio César Cazella and Cláudia de Souza Libânio from the Federal University of Health Sciences of Porto Alegre (UFCSPA) for inspiring us to write this review.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
SPECTsingle photon emission tomography
PETpositron emission tomography
HEUhighly enriched uranium
VHLvirtual health library
A EOB activity at the end of the bombardment
TTYthick target yield
SYsaturation yield
LINACSlinear particle accelerators
NPsnanoparticles
LWFAlaser wakefield accelerator
TNSAtarget normal sheath acceleration

Appendix A

Table A1. Intended radioisotope production summary.
Table A1. Intended radioisotope production summary.
Intended
Radioisotope
Half-Life
[h]
DecayDaughter
Isotope
TechnologyAccelerated
Particle
Production RouteReference
1 64 Cu12.7 β (39%) 64 ZnCyclotronProton 64 Ni(p,n) 64 CuThisgaard et al. [35]
β + (61%)
120 m Sb138.2 β + 120 Sn nat Sn(p,x) 120 m Sb
122 Sb65.4 β 122 Te nat Sn(p,x) 122 Sb
2 18 F1.83 β + 18 OCyclotronProton 18 O(p,n) 18 FRoeda et al. [36]
3 131 I192.5 β 131 XeResearch
reactor
Neutron 130 Te(n, γ ) 131 Te 131 IAchoribo et al. [37]
4 94 m Tc0.867 β + 94 MoCyclotronProton 94 Mo(p,n) 94 m TcHoehr et al. [38]
5 99 Mo66.0 β 99 m TcCyclotronProton 232 Th(p,f) 99 MoAbbas et al. [39]
6 13 N0.166 β + 13 CPlasma
focus device
Deuteron 13 C(d,n) 13 NShirani et al. [40]
7 161 Ho2.48Ec 161 DyCyclotronProton nat Dy(p,xn) 161 HoTárkányi et al. [41]
8 67 Ga78.3Ec 67 ZnCyclotronProton 68 Zn(p,2n) 67 GaMartins et al. [42]
9 169 Er225.4 β 169 TmResearch
reactor
Neutron 168 Er(n, γ ) 169 ErChakravarty et al. [43]
10 99 m Tc6.01 γ 99 TcCyclotronProton 100 Mo(p,2n) 99 m TcBénard et al. [44]
11 44 Sc4.04 β + 44 CaCyclotronProton 44 Ca(p,n) 44 ScHoehr et al. [45]
12 63 Zn0.641 β + 63 CuCyclotronProton 63 Cu(p,n) 63 ZnDeGrado et al. [46]
13 103 Ag1.10 β + 103 PdCyclotron 3 He nat Pd( 3 He,pxn) 103 AgAl-Abyad et al. [47]
104 Ag1.16 β + 104 Pd nat Pd( 3 He,pxn) 104 Ag
105 Ag991.0 β + 105 Pd nat Pd( 3 He,pxn) 105 Ag
106 m Ag198.7 β + 106 Cd nat Pd( 3 He,pxn) 106 m Ag
111 Ag178.8 β 111 Cd nat Pd( 3 He,pxn) 111 Ag
112 Ag3.13 β 112 Cd nat Pd( 3 He,pxn) 112 Ag
104 Cd0.962 β + 104 Ag nat Pd( 3 He,xn) 104 Cd
105 Cd0.925 β + 105 Ag nat Pd( 3 He,xn) 105 Cd
107 Cd6.50 β + 107 m Ag nat Pd( 3 He,xn) 107 Cd
111 m Cd0.808 β + 111 Cd nat Pd( 3 He,xn) 111 m Cd
14 211 At7.21 α 207 BiCyclotronProton 209 Bi( α ,2n) 211 AtMartin et al. [48]
15 89 Zr78.4 β + 89 YCyclotronProton 89 Y(p,n) 89 ZrSiikanen et al. [49]
16 44 Sc4.04 β + 44 CaCyclotronProton 44 Ca(p,n) 44 ScValdovinos et al. [50]
17 99 m Tc6.01 γ 99 TcCyclotronProton 100 Mo(p,2n) 99 m TcSchaffer et al. [51]
18 86 Y14.7 β + 86 SrCyclotronProton nat Sr(p,n) 86 YOehlke et al. [52]
89 Zr78.4 β + 89 Y 89 Y(p,n) 89 Zr
68 Ga1.13 β + 68 Zn nat Zn(p,x) 68 Ga
19 169 Yb1024Ec 169 TmResearch
reactor
Neutron nat Yb 2 O 3 (n, γ ) 169 YbSaxena et al. [53]
20 11 C0.339 β + 11 BCyclotronProton 14 N(p, α ) 11 CMoon et al. [54]
21 44 Sc4.04 β + 44 CaCyclotronProton 44 Ca(p,n) 44 Scvan der Meulen et al. [55]
22 44 Sc4.04 β + 44 CaCyclotronDeuteron 44 Ca(d,2n) 44 ScDuchemin et al. [56]
44 m Sc58.6 γ 44 Sc 44 Ca(d,2n) 44 m Sc
23 61 Cu3.33 β + 61 NiCyclotronProton nat Zn(p, α ) 61 CuAsad et al. [57]
64 Zn(p, α ) 61 Cu
24 161 Tb165.6 β 161 DyResearch
reactor
Neutron 160 Gd(n, γ ) 161 Gd 161 TbAziz et al. [58]
25 186 Re89.0 β 186 OsCyclotronProton nat W(p,n) 186 ReKakavand et al. [59]
26 43 Sc3.89 β + 43 CaCyclotron α nat Ca( α ,p) 43 ScSzkliniarz et al. [60]
40 Ca( α ,p) 43 Sc
44 Sc4.04 β + 44 Ca 42 Ca( α ,np+pn) 44 Sc
44 m Sc58.6 γ 44 Sc 42 Ca( α ,np+pn) 44 m Sc
27 66 Ga9.49 β + 66 ZnTandetronProton nat Zn(p,x) 66 GaFraile et al. [61]
68 Ga1.13 β + 68 Zn nat Zn(p,x) 68 Ga
28 66 Ga9.49 β + 66 ZnCyclotronProton 66 Zn(p,n) 66 GaCho et al. [62]
29 61 Cu3.33 β + 61 Ni Cyclotron Proton nat Zn(p, α ) 61 Cu do Carmo et al. [63]
66 Ga9.49 β + 66 Zn nat Zn(p,x) 66 Ga
67 Ga78.2Ec 67 Zn nat Zn(p,x) 67 Ga
68 Ga1.13 β + 68 Zn nat Zn(p,n) 68 Ga
65 Zn5857 β + 65 Cu nat Zn(p,pn) 65 Zn
30 47 Sc80.4 β 47 TiCyclotronProton nat Ca(p,2n) 64 CuMisiak et al. [64]
31 64 Cu12.7 β (39%) 64 ZnCyclotronProton 64 Ni(p,n) 64 CuXie et al. [65]
β + (61%)
32 43 Sc3.89 β + 43 CaCyclotronDeuteron 42 Ca(d,n) 43 ScSitarz et al. [66]
43 Ca(p,n) 43 Sc
44 Sc4.04 β + 44 Ca nat Ca(p,n) 44 Sc
Proton 44 Ca(p,n) 44 Sc
47 Sc80.4 β 47 Ti 48 Ca(p,2n) 47 Sc
48 Ti(p,2p) 47 Sc
33 100 Rh20.8 β + 100 RuCyclotron 3 He 103 Rh( 3 He,x) 100 RhAli et al. [67]
101 m Rh104.2Ec (94%) 101 Ru 103 Rh( 3 He, α +n) 101 m Rh
γ (6%) 101 Rh
103 Ag1.10 β + 103 Pd 103 Rh( 3 He,3n) 103 Ag
104 m Ag0.558 β + 104 Pd 103 Rh( 3 He,2n) 104 m Ag
104 Ag1.15 β + 103 Rh( 3 He,2n) 104 Ag
103 Pd407.8Ec 103 Rh 103 Rh( 3 He,n) 103 Pd
34 65 Zn5857Ec 65 CuResearch
reactor
Neutron 64 Zn(n, γ ) 65 ZnKarimi et al. [68]
35 106 m Ag19.2 β + 106 PdCyclotron α nat Ag( α ,x) 106 m AgDitrói et al. [69]
36 15 O0.034 β + 15 NCyclotronDeuteron 15 N(d,n) 15 OIguchi et al. [70]
37 72 Se201.6Ec 72 AsLinear particle
accelerator
Proton 75 As(p,4n) 72 SeDeGraffenreid et al. [71]
38 186 Re89.2 β 186 OsResearch
reactor
Neutron 185 Re(n, γ ) 186 RePourhabib et al. [72]
188 Re17.0 β 188 Os 187 Re(n, γ ) 188 Re
39 67 Cu61.8 β + 67 ZnCyclotron 70 Zn 15 + H 2 ( 70 Zn 15 + , α +p) 67 CuSouliotis et al. [73]
40 225 Ac240.0 α 221 FrCyclotronProton 232 Th(p,x) 225 AcRobertson et al. [74]
225 Ra357.6 β 225 Ac 232 Th(n,x) 225 Ra
41 94 Tc4.89 β + 94 MoCyclotronProton nat Mo(p,x) 94 TcAhmed et al. [75]
95 Tc20.0 β + 95 Mo nat Mo(p,x) 95 Tc
95 m Tc1464 β + 95 Mo nat Mo(p,x) 95 m Tc
96 Tc102.7 β + 96 Mo nat Mo(p,x) 96 Tc
99 m Tc6.01 γ 99 Tc nat Mo(p,x) 99 m Tc
99 Mo66.0 β 99 m Tc nat Mo(p,x) 99 Mo
42 44 Sc4.04 β + 44 CaCyclotronProton 44 Ca(p,n) 44 ScAlnahwi et al. [76]
43 99 m Tc6.01 γ 99 TcCyclotronProton nat Mo(p,x) 99 m TcKambali et al. [77]
99 Mo66.0 β 99 m Tc 98 Mo(p,x) 99 Mo
44 44 Sc4.04 β + 44 CaCyclotronProton 44 Ca(p,n) 44 Scvan der Meulen et al. [78]
45 13 N0.165 β + 13 CCyclotronProton 16 O(p, α ) 13 NAbel et al. [79]
18 F1.82 β + 18 O 18 O(p,n) 18 F
46 167 Tm222.0Ec 167 ErCyclotronProton nat Er(p,x) 167 TmHeinke et al. [80]
47 155 Tb127.7Ec 155 GdCyclotronProton 155 Gd(p,n) 155 TbFavaretto et al. [81]
156 Gd(p,2n) 155 Tb
48 89 Zr78.4 β + 89 YCyclotronProton 89 Y(p,n) 89 ZtCisterino et al. [82]
Table A2. Cyclotron produced radioisotopes.
Table A2. Cyclotron produced radioisotopes.
Production RouteKinetic Energy [MeV]Current [ μ A]Irradiation Time [h] A EOB [kBq]Reference
1 64 Ni(p,n) 64 Cu16.11211.26 8.2 × 10 6 Thisgaard et al. [35]
nat Sn(p,x) 120 m Sb150 3.71 × 10 4
nat Sn(p,x) 122 Sb 3.30 × 10 4
2 18 O(p,n) 18 F18200.5 2.96 × 10 7 Roeda et al. [36]
4 94 Mo(p,n) 94 m Tc1351 1.1 × 10 5 Hoehr et al. [38]
5 232 Th(p,f) 99 Mo29.511 3.7 × 10 3 Abbas et al. [39]
26.5 3.4 × 10 3
7 nat Dy(p,xn) 161 Ho36611.18 1.89 × 10 6 Tárkányi et al. [41]
8 68 Zn(p,2n) 67 Ga26101 3.9 × 10 5 Martins et al. [42]
10 100 Mo(p,2n) 99 m Tc182406.9 3.48 × 10 8 Bénard et al. [44]
11 44 Ca(p,n) 44 Sc137.61 5.5 × 10 3 Hoehr et al. [45]
12 63 Cu(p,n) 63 Zn14201 1.84 × 10 6 DeGrado et al. [46]
13 nat Pd( 3 He,pxn) 103 Ag270.1501 5.3 × 10 2 Al-Abyad et al. [47]
nat Pd( 3 He,pxn) 104 Ag 6.83 × 10 3
nat Pd( 3 He,pxn) 105 Ag30
nat Pd( 3 He,pxn) 106 m Ag35.9
nat Pd( 3 He,pxn) 111 Ag26.9
nat Pd( 3 He,pxn) 112 Ag102
nat Pd( 3 He,xn) 104 Cd 1.48 × 10 3
nat Pd( 3 He,xn) 105 Cd 5.95 × 10 3
nat Pd( 3 He,xn) 107 Cd690
nat Pd( 3 He,xn) 111 m Cd466
14 209 Bi( α ,2n) 211 At27.80.3274 3.91 × 10 4 Martin et al. [48]
25.30.192 7.91 × 10 3
15 89 Y(p,n) 89 Zr12.8453 6.3 × 10 6 Siikanen et al. [49]
16 44 Ca(p,n) 44 Sc15.6251 4.11 × 10 6 Valdovinos et al. [50]
17 100 Mo(p,2n) 99 m Tc16.51306 1.74 × 10 8 Schaffer et al. [51]
18 nat Sr(p,n) 86 Y134.51 7.4 × 10 3 Oehlke et al. [52]
89 Y(p,n) 89 Zr7 3.2 × 10 4
nat Zn(p,x) 68 Ga 4.8 × 10 5
20 14 N(p, α ) 11 C13600.5 8.66 × 10 7 Moon et al. [54]
21 44 Ca(p,n) 44 Sc11501.5 1.9 × 10 6 van der Meulen et al. [55]
22 44 Ca(d,2n) 44 Sc300.0540.5 7.5 × 10 4 Duchemin et al. [56]
44 Ca(d,2n) 44 m Sc 1.9 × 10 3
23 nat Zn(p, α ) 61 Cu11.7200.5 1.4 × 10 5 Asad et al. [57]
64 Zn(p, α ) 61 Cu40 6.2 × 10 5
25 nat W(p,n) 186 Re15205 5.19 × 10 4 Kakavand et al. [59]
26 nat Ca( α ,p) 43 Sc200.0501.5 5.5 × 10 3 Szkliniarz et al. [60]
40 Ca( α ,p) 43 Sc 5.8 × 10 3
42 Ca( α ,np+pn) 44 Sc29 2.1 × 10 3
42 Ca( α ,np+pn) 44 m Sc250
28 66 Zn(p,n) 66 Ga14.250.083508Cho et al. [62]
8195
29 nat Zn(p, α ) 61 Cu16.9300.75 2.72 × 10 5 do Carmo et al. [63]
nat Zn(p,x) 66 Ga 3.45 × 10 5
nat Zn(p,x) 67 Ga 2.24 × 10 4
nat Zn(p,n) 68 Ga 2.21 × 10 6
nat Zn(p,pn) 65 Zn225
30 nat Ca(p,2n) 47 Sc20.50.03521.1Misiak et al. [64]
31 64 Ni(p,n) 64 Cu12.5205-7 7.40 × 10 6 Xie et al. [65]
42 Ca(d,n) 43 Sc6.84 1.29 × 10 5
32 43 Ca(p,n) 43 Sc15.2 1 8 9.1 × 10 5 Sitarz et al. [66]
nat Ca(p,n) 44 Sc 5.0 × 10 4
44 Ca(p,n) 44 Sc 2.24 × 10 6
48 Ca(p,2n) 47 Sc22.8 4.2 × 10 5
48 Ti(p,2p) 47 Sc28 2.0 × 10 5
33 103 Rh( 3 He,x) 100 Rh270.11118Ali et al. [67]
103 Rh( 3 He, α +n) 101 m Rh30
103 Rh( 3 He,3n) 103 Ag 2.88 × 10 4
103 Rh( 3 He,2n) 104 m Ag 1.13 × 10 4
103 Rh( 3 He,2n) 104 Ag 8.87 × 10 3
103 Rh( 3 He,n) 103 Pd270
35 nat Ag( α ,x) 106 m Ag37.621238Ditrói et al. [69]
36 14 N(d,n) 15 O3.5400.017 2.22 × 10 6 Iguchi et al. [70]
39H 2 ( 70 Zn 15 + , α +p) 67 Cu150.0086.51.6Souliotis et al. [73]
40 232 Th(p,x) 225 Ac4807236.5 5.24 × 10 5 Robertson et al. [74]
232 Th(p,x) 225 Ra 8.6 × 10 4
41 nat Mo(p,x) 94 Tc260.0655 2.00 × 10 4 Ahmed et al. [75]
nat Mo(p,x) 95 Tc 1.55 × 10 4
nat Mo(p,x) 95 m Tc86.9
nat Mo(p,x) 96 Tc 2.42 × 10 3
nat Mo(p,x) 99 m Tc 7.14 × 10 4
nat Mo(p,x) 99 Mo 3.19 × 10 3
42 68 Zn(p,n) 68 Ga13351.5 1.44 × 10 8 Alnahwi et al. [76]
43 nat Mo(n,x) 99 m Tc11200.08339Kambali et al. [77]
98 Mo(n,x) 99 Mo33
44 44 Ca(p,n) 44 Sc1818.45.75 3.21 × 10 6 van der Meulen et al. [78]
45 16 O(p, α ) 13 N165.6–343.3 9.25 × 10 6 Abel et al. [79]
18 O(p,n) 18 F 2.15 × 10 5
46 nat Er(p,x) 167 Tm22.8508 2.61 × 10 5 Heinke et al. [80]
47 155 Gd(p,n) 155 Tb10508 2.0 × 10 5 Favaretto et al. [81]
156 Gd(p,2n) 155 Tb23 1.7 × 10 6
48 89 Y(p,x) 89 Zr12505 3.45 × 10 6 Cisterino et al. [82]
Table A3. Research reactor-produced radioisotopes.
Table A3. Research reactor-produced radioisotopes.
Production RouteTarget Mass [mg]Neutron Flux [n/cm 2 s]Irradiation
Time [h]
A EOB [kBq]Reference
3 130 Te(n, γ ) 131 Te 131 I5000 1.0 × 10 12 4400Achoribo et al. [37]
9 168 Er(n, γ ) 169 Er10 8.0 × 10 13 504 3.88 × 10 6 Chakravarty et al. [43]
19 nat Yb 2 O 3 (n, γ ) 169 Yb5 5.0 × 10 13 168 5.84 × 10 5 Saxena et al. [53]
24 160 Gd(n, γ ) 161 Gd 161 Tb100 1.0 × 10 14 108 3.92 × 10 6 Aziz et al. [58]
34 64 Zn(n, γ ) 65 Zn1000 4.5 × 10 13 0.5 6.16 × 10 3 Karimi et al. [68]
38 185 Re(n, γ ) 186 Re1 3.0 × 10 13 96 2.13 × 10 6 Pourhabib et al. [72]
187 Re(n, γ ) 188 Re 4.55 × 10 6
Table A4. Other technologies produced radioisotopes.
Table A4. Other technologies produced radioisotopes.
TechnologyProduction RouteKinetic Energy
[MeV]
Current
[ μ A]
Irradiation
Time [h]
A EOB [kBq]Reference
6Plasma
focus device
12 C(d,n) 13 N0.33not
applicable
0.0970.014Shirani et al. [40]
27Tandetron nat Zn(p,n) 66 Ga9 6.3 × 10 3 0.1068.5Fraile et al. [61]
nat Zn(p,n) 68 Ga99
37Linear particle
accelerator
75 As(p,4n) 72 Se49.5163.568.9 1.38 × 10 7 DeGraffenreid et al. [71]

References

  1. Yeong, C.H.; Cheng, M.H.; Ng, K.H. Therapeutic radionuclides in nuclear medicine: Current and future prospects. J. Zhejiang Univ. Sci. B 2014, 15, 845–863. [Google Scholar] [CrossRef] [Green Version]
  2. Crestoni, M.E. Radiopharmaceuticals for diagnosis and therapy. In Reference Module in Chemistry, Molecular Sciences and Chemical Engineering; Elsevier: Amsterdam, The Netherlands, 2018. [Google Scholar] [CrossRef]
  3. Radioisotopes in Medicine. Available online: https://world-nuclear.org/information-library/non-power-nuclear-applications/radioisotopes-research/radioisotopes-in-medicine.aspx (accessed on 23 August 2022).
  4. Qaim, S.M. Therapeutic radionuclides and nuclear data. Radiochim. Acta 2001, 89, 297–304. [Google Scholar] [CrossRef] [Green Version]
  5. Mallard, J.R.; Myers, M.J. Clinical Applications of a Gamma Camera. Phys. Med. Biol. 1963, 8, 183–192. [Google Scholar] [CrossRef] [PubMed]
  6. Subramanian, G.; Mcafee, J.G. A New Complex of 99mTc for Skeletal Imaging. Radiology 1971, 99, 192–196. [Google Scholar] [CrossRef] [PubMed]
  7. Bisi, G.; Sciagrà, R.; Santoro, G.M.; Cerisano, G.; Vella, A.; Zerauschek, F.; Fazzini, P.F. Myocardial scintigraphy with Tc-99m-teboroxime: Its feasibility and the evaluation of its diagnostic reliability. A comparison with thallium-201 and coronary angiography. G Ital. Cardiol. 1992, 22, 795–805. [Google Scholar]
  8. De Klerk, J.M.; van Rijk, P.P.; van Dongen, A.J.; Deenstra, M.; Bànki, J.H.; van het Schup, A.D. Can technetium 99 m bisdiethylphosphinoethanebis-t butylisocyanide (99mTc-DEPIC) be used for routine radionuclide ventriculography? Eur. J. Nucl. Med. 1991, 18, 317–320. [Google Scholar] [CrossRef]
  9. Ludwing, C.; Chicherio, C.; Terraneo, L.; Magistretti, P.; de Ribaupierre, A.; Slosman, D. Functional imaging studies of cognition using 99mTc-HMPAO SPECT: Empirical validation using the n-back working memory paradigm. Eur. J. Nucl. Med. Mol. Imaging 2008, 35, 695–703. [Google Scholar] [CrossRef] [Green Version]
  10. Seok, J.W.; Choi, Y.S.; Chong, S.; Kwon, G.Y.; Chung, Y.J.; Kum, B.G.; Park, S.J. Sentinel lymph node identification with radiopharmaceuticals in patients with breast cancer: A comparison of 99mTc-tin colloid and 99mTc-phytate efficiency. Breast Cancer Res. Treat. 2010, 122, 453–457. [Google Scholar] [CrossRef]
  11. Willkomm, P.; Bender, H.; Bangard, M.; Decker, P.; Grünwald, F.; Biersack, H.J. FDG PET and immunoscintigraphy with 99mTc-labeled antibody fragments for detection of the recurrence of colorectal carcinoma. J. Nucl. Med. 2000, 41, 1657–1663. [Google Scholar]
  12. Popescu, H.I.; Lessem, J.; Erjavec, M.; Füger, G.F. In vivo labelling of RBC with 99mTc for blood pool imaging using different stannous radiopharmaceuticals. J. Nucl. Med. 1984, 9, 295–299. [Google Scholar] [CrossRef] [PubMed]
  13. Chakraborty, D.; Sunil, H.V.; Mittal, B.R.; Bhattacharya, A.; Singh, B.; Chawla, Y. Role of Tc99m sulfur colloid scintigraphy in differentiating non-cirrhotic portal fibrosis from cirrhosis liver. Indian J. Nucl. Med. 2010, 25, 139–142. [Google Scholar] [CrossRef] [PubMed]
  14. Diamond, R.H.; Rothstein, R.D.; Alavi, A. The role of cimetidine-enhanced technetium-99m-pertechnetate imaging for visualizing Meckel’s diverticulum. J. Nucl. Med. 1991, 32, 1422–1424. [Google Scholar] [PubMed]
  15. Kanoun, S.; Rossi, C.; Casasnovas, O. [18F]FDG-PET/CT in Hodgkin Lymphoma: Current Usefulness and Perspectives. Cancers 2018, 10, 145. [Google Scholar] [CrossRef] [Green Version]
  16. Almuhaideb, A.; Papathanasiou, N.; Bomanji, J. 18F-FDG PET/CT imaging in oncology. Ann. Saudi Med. 2011, 31, 3–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Reed, E.; Lutsenko, S.; Bandmann, O. Animal models of Wilson disease. J. Neurochem. 2018, 146, 356–373. [Google Scholar] [CrossRef] [Green Version]
  18. Hancock, C.N.; Stockwin, L.H.; Han, B.; Divelbuss, R.D.; Jun, J.H.; Malhotra, S.V.; Hollingshead, M.G.; Newton, D.L. A copper chelate of thiosemicarbazone NSC 689534 induces oxidative/ER stress and inhibits tumor growth in vitro and in vivo. Free Radic. Biol. Med. 2011, 50, 110–121. [Google Scholar] [CrossRef] [Green Version]
  19. Parry, J.J.; Andrews, R.; Rogers, B.E. MicroPET imaging of breast cancer using radiolabeled bombesin analogs targeting the gastrin-releasing peptide receptor. Breast Cancer Res. Treat. 2007, 101, 175–183. [Google Scholar] [CrossRef]
  20. Lewis, S.J.; Laforest, R.; Buettner, T.L.; Song, S.-K.; Fujibayashi, Y.; Connett, J.M.; Welch, M.J. Copper-64-diacetyl-bis(N4-methylthiosemicarbazone): An agent for radiotherapy. Proc. Natl. Acad. Sci. USA 2001, 98, 1206–1211. [Google Scholar] [CrossRef] [Green Version]
  21. Radioisotope Production for Nuclear Medicine. Available online: https://www.radioactivity.eu.com/site/pages/Radioisotope_Production.htm (accessed on 23 August 2022).
  22. Qaim, S.M. Comparison of reactor and cyclotron production of medically important radioisotopes, with special reference to 99Mo/99mTc, 64,67Cu and 103Pd. In Utilization Related Design Features of Research Reactors: A Compendium; International Atomic Energy Agency (IAEA): Vienna, Austria, 2007; pp. 135–143. [Google Scholar]
  23. International Atomic Energy Agency. The Applications of Research Reactors. Report of An Advisory Group Meeting, Vienna. 2001, p. 71. Available online: https://inis.iaea.org/search/search.aspx?orig_q=RN:32048858 (accessed on 25 August 2022).
  24. Bibler, N.E.; Coleman, C.J.; Kinard, W.F. Relative Yields of U-235 Fission Products Measured in a High Level Radioactive Sludge at Savannah River Site. Westinghouse Savannah River Company. 1992. Available online: https://digital.library.unt.edu/ark:/67531/metadc1194158 (accessed on 25 August 2022).
  25. Storage and Disposal of Radioactive Waste. Available online: https://world-nuclear.org/information-library/nuclear-fuel-cycle/nuclear-waste/storage-and-disposal-of-radioactive-waste.aspx (accessed on 23 August 2022).
  26. Cho, W.-K. Radiation risk assessment for the transport of radioisotopes using KRI-BGM B(U) type container. In IRPA12: 12. Congress of the International Radiation Protection Association: Strengthening Radiation Protection Worldwide, Highlights, Global Perspective and Future Trends; International Atomic Energy Agency (IAEA): Vienna, Austria, 2010. [Google Scholar]
  27. United States Environmental Protection Agency. Transportation of Radioactive Material. Available online: https://www.epa.gov/radtown/transportation-radioactive-material (accessed on 7 November 2022).
  28. Adedapo, K.S.; Onimode, Y.A.; Ejeh, J.E.; Adepoju, A.O. Avoidable challenges of a nuclear medicine facility in a developing nation. Indian J. Nucl. Med. 2013, 28, 195–199. [Google Scholar] [CrossRef]
  29. Van Noorden, R. Radioisotopes: The medical testing crisis. Nature 2013, 504, 202–204. [Google Scholar] [CrossRef] [Green Version]
  30. National Research Council (US) Committee on Medical Isotope Production without Highly Enriched Uranium. Medical Isotope Production without Highly Enriched Uranium; National Academies Press: Washington, DC, USA, 2009. [Google Scholar]
  31. Whittemore, R.; Knafl, K. The integrative review: Updated methodology. J. Adv. Nurs. 2005, 52, 546–553. [Google Scholar] [CrossRef] [PubMed]
  32. Available online: https://pesquisa.bvsalud.org/portal/advanced/?lang=en (accessed on 25 August 2022).
  33. Available online: https://www.sciencedirect.com/search (accessed on 25 August 2022).
  34. Available online: https://www.scopus.com/ (accessed on 25 August 2022).
  35. Thisgaard, H.; Jensen, M.; Elema, D.R. Medium to large scale radioisotope production for targeted radiotherapy using a small PET cyclotron. Appl. Radiat. Isot. 2011, 69, 1–7. [Google Scholar] [CrossRef] [PubMed]
  36. Roeda, D.; Kuhnast, B.; Damont, A.; Dollé, F. Synthesis of fluorine-18-labeled TSPO ligands for imaging neuroinflammation with Positron Emission Tomography. J. Fluor. Chem. 2012, 134, 107–114. [Google Scholar] [CrossRef]
  37. Elom Achoribo, A.S.; Akaho, E.H.; Nyarko, B.J.; Osae Shiloh, K.D.; Odame Duodu, G.; Gibrilla, A. Feasibility study for production of I-131 radioisotope using MNSR research reactor. Appl. Radiat. Isot. 2012, 70, 76–80. [Google Scholar] [CrossRef]
  38. Hoehr, C.; Morley, T.; Buckley, K.; Trinczek, M.; Hanemaayer, V.; Schaffer, P.; Ruth, T.; Bénard, F. Radiometals from liquid targets: 94mTc production using a standard water target on a 13 MeV cyclotron. Appl. Radiat. Isot. 2012, 70, 2308–2312. [Google Scholar] [CrossRef]
  39. Abbas, K.; Holzwarth, U.; Simonelli, F.; Kozempel, J.; Cydzik, I.; Bulgheroni, A.; Cotogno, G.; Apostolidis, C.; Bruchertseifer, F.; Morgenstern, A. Feasibility of 99Mo production by proton-induced fission of 232Th. Nucl. Instrum. Methods Phys. Res. B Nucl. Instrum. Meth. B 2012, 278, 20–25. [Google Scholar] [CrossRef]
  40. Shirani, B.; Abbasi, F.; Nikbakht, M. Production of 13N by 12C(d,n)13N reaction in a medium energy plasma focus. Appl. Radiat. Isot. 2013, 74, 86–90. [Google Scholar] [CrossRef]
  41. Tárkányi, F.; Ditroi, F.; Hermanne, A.; Takács, S.; Ignatyuk, A. Investigation of production routes for the 161Ho Auger-electron emitting radiolanthanide, a candidate for therapy. J. Radioanal. Nucl. Chem. 2013, 298, 277–286. [Google Scholar] [CrossRef] [Green Version]
  42. Martins, P.d.A.; Osso Junior, J.A. Thermal diffusion of 67Ga from irradiated Zn targets. Appl. Radiat. Isot. 2013, 82, 279–282. Available online: http://repositorio.ipen.br/handle/123456789/4005 (accessed on 25 August 2022). [CrossRef]
  43. Chakravarty, R.; Chakraborty, S.; Chirayil, V.; Dash, A. Reactor production and electrochemical purification of 169Er: A potential step forward for its utilization in in vivo therapeutic applications. Nucl. Med. Biol. 2014, 41, 163–170. [Google Scholar] [CrossRef]
  44. Bénard, F.; Buckley, K.R.; Ruth, T.J.; Zeisler, S.K.; Klug, J.; Hanemaayer, V.; Vuckovic, M.; Hou, X.; Celler, A.; Appiah, J.P.; et al. Implementation of Multi-Curie Production of 99mTc by Conventional Medical Cyclotrons. J. Nucl. Med. 2014, 55, 1017–1022. [Google Scholar] [CrossRef] [PubMed]
  45. Hoehr, C.; Oehlke, E.; Benard, F.; Lee, C.J.; Hou, X.; Badesso, B.; Ferguson, S.; Miao, Q.; Yang, H.; Buckley, K.; et al. 44gSc production using a water target on a 13 MeV cyclotron. Nucl. Med. Biol. 2014, 41, 401–406. [Google Scholar] [CrossRef] [PubMed]
  46. DeGrado, T.R.; Pandey, M.K.; Byrne, J.F.; Engelbrecht, H.P.; Jiang, H.; Packard, A.B.; Thomas, K.A.; Jacobson, M.S.; Curran, G.L.; Lowe, V.J. Preparation and preliminary evaluation of 63Zn-zinc citrate as a novel PET imaging biomarker for zinc. J. Nucl. Med. 2014, 55, 1348–1354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Al-Abyad, M.; Tárkányi, F.; Ditrói, F.; Takács, S. Excitation function of 3He-particle induced nuclear reactions on natural palladium. Appl. Radiat. Isot. 2014, 94, 191–199. [Google Scholar] [CrossRef]
  48. Martin, T.M.; Bhakta, V.; Al-Harbi, A.; Hackemack, M.; Tabacaru, G.; Tribble, R.; Shankar, S.; Akabani, G. Preliminary production of 211At at the Texas A&M University Cyclotron Institute. Health Phys. 2014, 107, 1–9. [Google Scholar] [CrossRef]
  49. Siikanen, J.; Tran, T.A.; Olsson, T.G.; Strand, S.-E.; Sandell, A. A solid target system with remote handling of irradiated targets for PET cyclotrons. Appl. Radiat. Isot. 2014, 94, 294–301. [Google Scholar] [CrossRef]
  50. Valdovinos, H.F.; Hernandez, R.; Barnhart, T.E.; Graves, S.; Cai, W.; Nickles, R.J. Separation of cyclotron-produced 44Sc from a natural calcium target using a dipentyl pentylphosphonate functionalized extraction resin. Appl. Radiat. Isot. 2015, 95, 23–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Schaffer, P.; Benard, F.; Bernstein, A.; Buckley, K.; Celler, A.; Cockburn, N.; Corsaut, J.; Dodd, M.; Economou, C.; Eriksson, T.; et al. Direct Production of 99mTc via 100Mo(p,2n) on Small Medical Cyclotrons. Physics Procedia 2015, 66, 383–395. [Google Scholar] [CrossRef] [Green Version]
  52. Oehlke, E.; Hoehr, C.; Hou, X.; Hanemaayer, V.; Zeisler, S.; Adam, M.J.; Ruth, T.J.; Celler, A.; Buckley, K.; Benard, F.; et al. Production of Y-86 and other radiometals for research purposes using a solution target system. Nucl. Med. Biol. 2015, 42, 842–849. [Google Scholar] [CrossRef]
  53. Saxena, S.K.; Kumar, Y.; Jagadeesan, K.C.; Nuwad, J.; Bamankar, Y.R.; Dash, A. Studies on the development of 169Yb-brachytherapy seeds: New generation brachytherapy sources for the management of cancer. Appl. Radiat. Isot. 2015, 101, 75–82. [Google Scholar] [CrossRef]
  54. Moon, B.; Lee, H.; Lee, W.; Hur, M.G.; Yang, S.; Lee, B.; Kim, S. Development of additive [11C]CO2 target system in the KOTRON-13 cyclotron and its application for [11C]radiopharmaceutical production. Nucl. Instrum. Methods Phys. Res. B Nucl. Instrum. Meth. B 2015, 356, 1–7. [Google Scholar] [CrossRef]
  55. Van der Meulen, N.P.; Bunka, M.; Domnanich, K.A.; Müller, C.; Haller, S.; Vermeulen, C.; Türler, A.; Schibli, R. Cyclotron production of 44Sc: From bench to bedside. Nucl. Med. Biol. 2015, 42, 745–751. [Google Scholar] [CrossRef] [PubMed]
  56. Duchemin, C.; Guertin, A.; Haddad, F.; Michel, N.; Métivier, V. Production of scandium-44m and scandium-44g with deuterons on calcium-44: Cross section measurements and production yield calculations. Phys. Med. Biol. 2015, 60, 6847–6864. [Google Scholar] [CrossRef]
  57. Asad, A.H.; Smith, S.V.; Morandeau, L.M.; Chan, S.; Jeffery, C.M.; Price, R.I. Production of 61Cu by the natZn(p,α) reaction: Improved separation and specific activity determination by titration with three chelators. J. Radioanal. Nucl. Chem. 2015, 307, 899–906. [Google Scholar] [CrossRef]
  58. Aziz, A.; Artha, W. Radiochemical Separation of 161Tb from Gd/Tb Matrix Using Ln Resin Column. Indones. J. Chem. 2016, 16. [Google Scholar] [CrossRef]
  59. Kakavand, T.; Mirzaii, M.; Khaleghi, M.; Eslami, M. Optimization of sedimentation of tungsten on copper substrate for production of 168gRe via 168W(p,n) nuclear reaction: Feasibility of using high current, long irradiation. Nucl. Instrum. Methods Phys. Res. Sect. B 2016, 366, 224–226. [Google Scholar] [CrossRef]
  60. Szkliniarz, K.; Sitarz, M.; Walczak, R.; Jastrzebski, J.; Bilewicz, A.; Choinski, J.; Jakubowski, A.; Majkowska, A.; Stolarz, A.; Trzcinska, A.; et al. Production of medical Sc radioisotopes with an alpha particle beam. Appl. Radiat. Isot. 2016, 118, 182–189. [Google Scholar] [CrossRef] [Green Version]
  61. Fraile, L.; Lopez Herraiz, J.; Udías, J.; Cal-Gonzalez, J.; García Corzo, P.M.; España, S.; Herranz, E.; Pérez-Liva, M.; Picado, E.; Vicente, E.; et al. Experimental validation of gallium production and isotope-dependent positron range correction in PET. Nucl. Instrum. Methods Phys. Res. Sect. B Nucl. Instrum. Methods A 2016, 814, 110–116. [Google Scholar] [CrossRef] [Green Version]
  62. Cho, J.; Wang, M.; Gonzalez-Lepera, C.; Mawlawi, O.; Cho, S.H. Development of bimetallic (Zn@Au) nanoparticles as potential PET-imageable radiosensitizers. Med. Phys. 2016, 43. [Google Scholar] [CrossRef] [Green Version]
  63. Do Carmo, S.J.C.; Alves, V.H.P.; Alves, F.; Abrunhosa, A.J. Fast and cost-effective cyclotron production of 61Cu using a natZn liquid target: An opportunity for radiopharmaceutical production and R&D. Dalton Trans. 2017, 46, 14556–14560. [Google Scholar] [CrossRef]
  64. Misiak, R.; Walczak, R.; Wąs, B.; Bartyzel, M.; Mietelski, J.W.; Bilewicz, A. 47Sc production development by cyclotron irradiation of 48Ca. J. Radioanal. Nucl. Chem. 2017, 313, 429–434. [Google Scholar] [CrossRef]
  65. Xie, Q.; Zhu, H.; Wang, F.; Meng, X.; Ren, Q.; Xia, C.; Yang, Z. Establishing Reliable Cu-64 Production Process: From Target Plating to Molecular Specific Tumor Micro-PET Imaging. Molecules 2017, 22, 641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Sitarz, M.; Szkliniarz, K.; Jastrzębski, J.; Choiński, J.; Guertin, A.; Haddad, F.; Jakubowski, A.; Kapinos, K.; Kisieliński, M.; Majkowska, A.; et al. Production of Sc medical radioisotopes with proton and deuteron beams. Appl. Radiat. Isot. 2018, 142, 104–112. [Google Scholar] [CrossRef] [PubMed]
  67. Ali, B.M.; Al-Abyad, M.; Kandil, S.; Solieman, A.H.M.; Ditrói, F. Excitation functions of 3He-particle-induced nuclear reactions on 103Rh: Experimental and theoretical investigations. Eur. Phys. J. Plus 2018, 133, 9. [Google Scholar] [CrossRef] [Green Version]
  68. Karimi, Z.; Sadeghi, M.; Rostampour, M. Assessment and estimation of 65Zn production yield via neutron induced reaction on natZnO and natZnONPs. Appl. Radiat. Isot. 2018, 141, 118–121. [Google Scholar] [CrossRef] [PubMed]
  69. Ditrói, F.; Takács, S.; Haba, H.; Komori, Y.; Aikawa, M.; Saito, M.; Murata, T. Investigation of alpha particle induced reactions on natural silver in the 40–50 MeV energy range. Nucl. Instruments Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2018, 436, 119–129. [Google Scholar] [CrossRef]
  70. Iguchi, S.; Moriguchi, T.; Yamazaki, M.; Hori, Y.; Koshino, K.; Toyoda, K.; Teuho, J.; Shimochi, S.; Terakawa, Y.; Fukuda, T.; et al. System evaluation of automated production and inhalation of 15O-labeled gaseous radiopharmaceuticals for the rapid 15O-oxygen PET examinations. EJNMMI Phys. 2018, 5, 37. [Google Scholar] [CrossRef]
  71. Degraffenreid, A.; Medvedev, D.; Phelps, T.; Gott, M.; Smith, S.; Jurisson, S.; Cutler, C. Cross-section measurements and production of 72Se with medium to high energy protons using arsenic containing targets. Radiochim. Acta. 2019, 107, 279–287. [Google Scholar] [CrossRef]
  72. Pourhabib, Z.; Ranjbar, H.; Bahrami Samani, A.; Shokri, A.A. Experimental and theoretical study of rhenium radioisotopes production for manufacturing of new compositional radiopharmaceuticals. Appl. Radiat. Isot. 2019, 145, 176–179. [Google Scholar] [CrossRef]
  73. Souliotis, G.A.; Rodrigues, M.R.D.; Wang, K.; Iacob, V.E.; Nica, N.; Roeder, B.; Tabacaru, G.; Yu, M.; Zanotti-Fregonara, P.; Bonasera, A. A novel approach to medical radioisotope production using inverse kinematics: A successful production test of the theranostic radionuclide 67Cu. Appl. Radiat. Isot. 2019, 149, 89–95. [Google Scholar] [CrossRef] [Green Version]
  74. Robertson, A.K.H.; Lobbezoo, A.; Moskven, L.; Schaffer, P.; Hoehr, C. Design of a Thorium Metal Target for 225Ac Production at TRIUMF. Instruments 2019, 3, 18. [Google Scholar] [CrossRef]
  75. Ahmed, A.A.; Wrońska, A.; Magiera, A.; Bartyzel, M.; Mietelski, J.W.; Misiak, R.; Wąs, B. Reexamination of Proton-induced Reactions on natMo at 19–26 MeV and Study of Target Yield of Resultant Radionuclides. Acta Phys. Pol. B 2019, 50, 1583. [Google Scholar] [CrossRef]
  76. Alnahwi, A.H.; Samia Ait-Mohand, S.T.; Beaudoin, J.-F.; Guérin, B. Automated radiosynthesis of 68Ga for large-scale routine production using 68Zn pressed target. Appl. Radiat. Isot. 2020, 156, 109014. [Google Scholar] [CrossRef] [PubMed]
  77. Kambali, I.; Rajiman, P.; Marlina, K.; Huda, N.; Listiawadi, F.D.; Astarina, H.; Ismuha, R.R.; Heranudin, S. Hari Spectral analysis of proton-irradiated natural MoO3 relevant for technetium-99 m radionuclide production. J. Math. Fundam. Sci. 2020, 52, 222–231. [Google Scholar] [CrossRef]
  78. Van der Meulen, N.P.; Hasler, R.; Talip, Z.; Grundler, P.V.; Favaretto, C.; Umbricht, C.A.; Müller, C.; Dellepiane, G.; Carzaniga, T.S.; Braccini, S. Developments toward the Implementation of 44Sc Production at a Medical Cyclotron. Molecules 2020, 25, 4706. [Google Scholar] [CrossRef] [PubMed]
  79. Abel, E.P.; Domnanich, K.; Kalman, C.; Walker, W.; Engle, J.W.; Barnhart, T.E.; Severin, G. Durability test of a flowing-water target for isotope harvesting. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2020, 478, 34–45. [Google Scholar] [CrossRef]
  80. Heinke, R.; Chevallay, E.; Chrysalidis, K.; Cocolios, T.E.; Duchemin, C.; Fedosseev, V.N.; Hurier, S.; Lambert, L.; Leenders, B.; Marsh, B.A.; et al. Efficient Production of High Specific Activity Thulium-167 at Paul Scherrer Institute and CERN-MEDICIS. Front. Med. 2021, 8, 712374. [Google Scholar] [CrossRef] [PubMed]
  81. Favaretto, C.; Talip, Z.; Borgna, F.; Grundler, P.V.; Dellepiane, G.; Sommerhalder, A.; Zhang, H.; Schibli, R.; Braccini, S.; Müller, C.; et al. Cyclotron production and radiochemical purification of terbium-155 for SPECT imaging. EJNMMI Radiopharm. Chem. 2021, 6, 37. [Google Scholar] [CrossRef]
  82. Cisternino, S.; Cazzola, E.; Skliarova, H.; Amico, J.; Malachini, M.; Gorgoni, G.; Anselmi-Tamburini, U.; Esposito, J. Target manufacturing by Spark Plasma Sintering for efficient 89Zr production. Nucl. Med. Biol. 2021, 104, 38–46. [Google Scholar] [CrossRef]
  83. Krasnov, N.N. Thick target yield. The International Appl. Radiat. Isot. 1974, 25, 223–227. [Google Scholar] [CrossRef]
  84. Brady, F.; Clark, J.C.; Luthra, S. Building on a 50-year legacy of the MRC Cyclotron Unit: The Hammersmith radiochemistry pioneering journey. J. Label. Compd. Radiopharm. 2007, 50, 903–926. [Google Scholar] [CrossRef]
  85. Peeva, A. Cyclotrons—What Are They and Where Can You Find Them. Available online: https://www.iaea.org/newscenter/news/cyclotrons-what-are-they-and-where-can-you-find-them (accessed on 25 August 2022).
  86. García-Toraño, E.; Pryrés, V.; Roteta, M.; Sánchez-Cabezudo, A.I.; Romero, E.; Ortega, A.M. Standardisation and precise determination of the half-life of 44Sc. Appl. Radiat. Isot. 2016, 109, 314–318. [Google Scholar] [CrossRef] [PubMed]
  87. Roesch, F. Scandium-44: Benefits of a long-lived PET radionuclide available from the (44)Ti/(44)Sc generator system. Curr. Radiopharm. 2012, 5, 187–201. [Google Scholar] [CrossRef] [PubMed]
  88. Rosar, F.; Buchholz, H.G.; Michels, S.; Hoffmann, M.A.; Piel, M.; Waldmann, C.M.; Rösch, F.; Reuss, S.; Schreckenberger, M. Image quality analysis of 44Sc on two preclinical PET scanners: A comparison to 68Ga. EJNMMI Phys. 2020, 16, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Kolsky, K.L.; Joshi, V.; Mausner, L.F.; Srivastava, S.C. Radiochemical purification of no-carrier-added scandium-47 for radioimmunotherapy. Appl. Radiat. Isot. 1998, 49, 1541–1549. [Google Scholar] [CrossRef]
  90. Müller, C.; Bunka, M.; Haller, S.; Köster, U.; Groehn, V.; Bernhardt, P.; van der Meuler, N.P.; Türler, A.; Schibli, R. Promising Prospects for 44Sc-/47Sc-Based Theragnostics: Application of 47Sc for Radionuclide Tumor Therapy in Mice. J. Nucl. Med. 2014, 55, 1658–1664. [Google Scholar] [CrossRef] [Green Version]
  91. Koumarianou, E.; Loktionova, N.S.; Fellner, M.; Roesch, F.; Thews, O.; Paelak, D.; Archumandritis, S.C.; Mikolajczak, R. 44Sc-DOTA-BN[2-14]NH2 in comparison to 68Ga-DOTA-BN[2-14]NH2 in pre-clinical investigation. Is 44Sc a potential radionuclide for PET? Appl. Radiat. Isot. 2012, 70, 2669–2676. [Google Scholar] [CrossRef]
  92. Radchenko, V.; Engle, J.W.; Medvedev, D.G.; Maassen, J.M.; Naranjo, C.M.; Unc, G.A.; Meyer, C.A.L.; Mastren, T.; Brugh, M.; Mausner, L.; et al. Proton-induced production and radiochemical isolation of 44Ti from scandium metal targets for 44Ti/44Sc generator development. Nucl. Med. Biol. 2017, 50, 25–32. [Google Scholar] [CrossRef] [Green Version]
  93. Oroujeni, M.; Xu, T.; Gagnon, K.; Rinne, S.S.; Weis, J.; Garousi, J.; Andersson, K.G.; Löfblom, J.; Orlova, A.; Tolmachev, V. The Use of a Non-Conventional Long-Lived Gallium Radioisotope 66Ga Improves Imaging Contrast of EGFR Expression in Malignant Tumours Using DFO-ZEGFR:2377 Affibody Molecule. Pharmaceutics 2021, 13, 292. [Google Scholar] [CrossRef]
  94. Goethals, P.; Coene, M.; Slegers, G.; Agon, P.; Deman, J.; Schelstraete, K. Cyclotron production of carrier-free 66Ga as a positron emitting label of albumin colloids for clinical use. Eur. J. Nucl. Med. 1988, 14, 152–154. [Google Scholar] [CrossRef]
  95. Engle, J.W.; Hong, H.; Zhang, Y.; Valdovinos, H.F.; Myklejord, D.V.; Barnhart, T.E.; Theuer, C.P.; Nickles, R.J.; Cai, W. Positron emission tomography imaging of tumor angiogenesis with a 66Ga-labeled monoclonal antibody. Mol. Pharm. 2012, 9, 1441–1448. [Google Scholar] [CrossRef] [PubMed]
  96. Ugur, Ö.; Kothari, P.J.; Finn, R.D.; Zanzonico, P.; Ruan, S.; Guenther, I.; Maecke, H.R.; Larson, S.M. Ga-66 labeled somatostatin analogue DOTA-DPhe1-Tyr3-octreotide as a potential agent for positron emission tomography imaging and receptor mediated internal radiotherapy of somatostatin receptor positive tumors. Nucl. Med. Biol. 2002, 29, 147–157. [Google Scholar] [CrossRef] [PubMed]
  97. Rinne, S.S.; Gagnon, K.; Abouzayed, A.; Tolnachev, V.; Orlova, A. 66Ga-PET-Imaging of GRPR-expression in prostate cancer: Production and characterization of [66Ga]Ga-NOTA-PEG2-RM26. Sci. Rep. 2021, 11, 3631. [Google Scholar] [CrossRef] [PubMed]
  98. Afshar-Oromieh, A.; Babich, J.W.; Kratochwil, C.; Giesel, F.L.; Eisenhu, M.; Kopka, K.; Haberkorn, U. The rise of PSMA ligands for diagnosis and therapy of prostate cancer. J. Nucl. Med. 2016, 57, 79S–89S. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Graham, M.M.; Gu, X.; Ginader, T.; Breheny, P.; Sunderland, J.J. 68Ga-DOTATOC Imaging of neuroendocrine tumors: A systematic review and metaanalysis. J. Nucl. Med. 2017, 58, 1452–1458. [Google Scholar] [CrossRef] [Green Version]
  100. Ehrhardt, G.J.; Welcj, M.J. A new germanium-63/gallium-68 generator. J. Nucl. Med. 1978, 19, 925–929. [Google Scholar]
  101. Breeman, W.A.; Verbruggen, A.M. The 68Ge/68Ga generator has highpotential, but when can we use 68Ga-labeled tracers in clinicalroutine? Eur. J. Nucl. Med. Mol. Imaging 2007, 34, 978–981. [Google Scholar] [CrossRef] [Green Version]
  102. Nayak, T.K.; Brechbiel, M.W. Radioimmunoimaging with longer-lived positron-emitting radionuclides: Potentials and challenges. Bioconjug. Chem. 2009, 20, 825–841. [Google Scholar] [CrossRef] [Green Version]
  103. Yoon, J.K.; Park, B.N.; Ryu, E.K.; An, Y.S.; Lee, S.J. Current perspectives on 89Zr-PET imaging. Int. J. Mol. Sci. 2020, 21, 4309. [Google Scholar] [CrossRef]
  104. Hovhannisyan, G.H.; Bakhshiyan, T.M.; Balabekyan, A.R.; Kerobyan, I.A. Production of 47Sc in photonuclear reactions on natTi targets at the bremsstrahlung endpoint energy of 30 and 40 MeV. Appl. Radiat. Isot. 2022, 182, 110138. [Google Scholar] [CrossRef]
  105. Sabel’nikov, A.V.; Maslov, O.D.; Molokanova, L.G.; Gustova, M.V.; Dmitriev, S.N. Preparation of 99Mo and 99mTc by 100Mo(γ,n) photonuclear reaction on an electron accelerator, MT-25 microtron. Radiochemistry 2006, 48, 191–194. [Google Scholar] [CrossRef]
  106. Vieira, N.D.; Maldonado, E.P.; Bonatto, A.; Nunes, R.P.; Banerjee, S.; Genezini, F.A.; Moralles, M.; Zuffi, A.V.F.; Samad, R.E. Laser wakefield electron accelerator: Possible use for radioisotope production. In Proceedings of the 2021 SBFoton International Optics and Photonics Conference (SBFoton IOPC), Sao Carlos, Brazil, 31 May–2 June 2021; pp. 1–6. [Google Scholar] [CrossRef]
  107. Tajima, T.; Dawson, J.M. Laser electron accelerator. Phys. Rev. Lett. 1979, 43, 267. [Google Scholar] [CrossRef] [Green Version]
  108. Esarey, E.; Schroeder, C.B.; Leemans, W.P. Physics of laser-driven plasma-based electron accelerator. Rev. Mod. Phys 2009, 81, 1229–1285. [Google Scholar] [CrossRef]
  109. Gonsalves, A.J.; Nakamura, K.; Daniels, J.; Benedetti, C.; Pieronek, C.; de Raadt, T.C.H.; Steinke, S.; Bin, J.H.; Bulanov, S.S.; van Tilborg, J.; et al. Petawatt Laser Guiding and Electron Beam Acceleration to 8 GeV in a Laser-Heated Capillary Discharge Waveguide. Phys. Rev. Lett. 2019, 122, 084801. [Google Scholar] [CrossRef] [Green Version]
  110. Rovige, L.; Huijts, J.; Andriyash, I.; Vernier, A.; Tomkus, V.; Girdauskas, V.; Raciukaitis, G.; Dudutis, J.; Stankevic, V.; Gecys, P.; et al. Demonstration of stable long-term operation of a kilohertz laser-plasma accelerator. Phys. Rev. Accel. Beams 2020, 23, 093401. [Google Scholar] [CrossRef]
  111. Tajima, T. Laser acceleration and its future. Proc. Jpn. Acad. Ser. B 2010, 86, 147–157. [Google Scholar] [CrossRef] [Green Version]
  112. Sun, Z. Review: Production of nuclear medicine radioisotopes with ultra-intense lasers. AIP Adv. 2021, 11, 040701. [Google Scholar] [CrossRef]
  113. Lobok, M.G.; Brantov, A.V.; Bychenkov, V.Y. Laser-based photonuclear production of medical isotopes and nuclear waste transmutation. Plasma Phys. Control. Fusion 2022, 64, 054002. [Google Scholar] [CrossRef]
  114. Shalloo, R.J.; Dann, S.J.D.; Gruse, J.-N.; Underwood, C.I.D.; Antoine, A.F.; Arran, C.; Backhouse, M.; Baird, C.D.; Balcazar, M.D.; Bourgeois, N.; et al. Automation and control of laser wakefield accelerators using Bayesian optimization. Nat. Commun. 2020, 11, 6355. [Google Scholar] [CrossRef]
  115. Jalas, S.; Kirchen, M.; Messner, P.; Winkler, P.; Hübner, L.; Dirkwinkel, J.; Schnepp, M.; Lehe, R.; Maier, A.R. Bayesian Optimization of a Laser-Plasma Accelerator. Phys. Rev. Lett. 2021, 126, 104801. [Google Scholar] [CrossRef]
  116. Snavely, R.A.; Key, M.H.; Hatchett, S.P.; Cowan, T.E.; Roth, M.; Phillips, T.W.; Stoyer, M.A.; Henry, E.A.; Sangster, T.C.; Singh, M.S.; et al. Intense High-Energy Proton Beams from Petawatt-Laser Irradiation of Solids. Phys. Rev. Lett. 2000, 85, 2945. [Google Scholar] [CrossRef] [PubMed]
  117. Hatchett, S.P.; Brown, C.G.; Cowan, T.E.; Henry, E.A.; Johnson, J.S.; Key, M.H.; Koch, J.A.; Langdon, A.B.; Lasinski, B.F.; Lee, R.W.; et al. Electron, photon, and ion beams from the relativistic interaction of Petawatt laser pulses with solid targets. Phys. Plasmas 2000, 7, 2076–2082. [Google Scholar] [CrossRef] [Green Version]
  118. Afshari, M.; Hornung, J.; Kleinschmidt, A.; Neumayer, P.; Bertini, D.; Bagnoud, V. Proton acceleration via the TNSA mechanism using a smoothed laser focus. AIP Adv. 2020, 10, 035023. [Google Scholar] [CrossRef] [Green Version]
  119. Perego, C.; Batani, D.; Zani, A.; Passoni, M. Target normal sheath acceleration analytical modeling, comparative study and developments. Rev. Sci. Instrum. 2012, 83, 02B502. [Google Scholar] [CrossRef] [PubMed]
  120. Clarke, R.; Dorkings, S.; Neely, D.; Musgrave, I. The production of patient dose level 99mTc medical radioisotope using laser-driven proton beams. In Proceedings of the Laser Acceleration of Electrons, Protons, and Ions II; and Medical Applications of Laser-Generated Beams of Particles II, and Harnessing Relativistic Plasma Waves III International Society for Optics and Photonics, Prague, Czech Republic, 15–18 April 2013; Volume 8779, p. 87791C. [Google Scholar]
  121. Romanov, D.V.; Bychenkov, V.Y.; Rozmus, W.; Capjack, C.E.; Fedosejevs, R. Self-organization of a plasma due to 3D evolution of the Weibel instability. Phys. Rev. Lett. 2004, 93, 215004. [Google Scholar] [CrossRef] [PubMed]
  122. Bychenkov, V.Y.; Brantov, A.V.; Mourou, G. Tc-99m production with ultrashort intense laser pulses. Laser Part. Beams 2014, 32, 605–611. [Google Scholar] [CrossRef]
Figure 1. Number of papers that utilized (a) the determined production technology and (b) accelerated particles to produce the intended radioisotope.
Figure 1. Number of papers that utilized (a) the determined production technology and (b) accelerated particles to produce the intended radioisotope.
Applsci 12 12569 g001
Figure 2. Number of papers that cited the production of each radioisotope.
Figure 2. Number of papers that cited the production of each radioisotope.
Applsci 12 12569 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nunes, B.S.; Rodrigues, E.R.F.; Fruscalso, J.A.P.; Nunes, R.P.; Bonatto, A.; Alva-Sánchez, M.S. Highly Enriched Uranium-Free Medical Radioisotope Production Methods: An Integrative Review. Appl. Sci. 2022, 12, 12569. https://doi.org/10.3390/app122412569

AMA Style

Nunes BS, Rodrigues ERF, Fruscalso JAP, Nunes RP, Bonatto A, Alva-Sánchez MS. Highly Enriched Uranium-Free Medical Radioisotope Production Methods: An Integrative Review. Applied Sciences. 2022; 12(24):12569. https://doi.org/10.3390/app122412569

Chicago/Turabian Style

Nunes, Bruno Silveira, Enio Rodrigo Fernandes Rodrigues, Jonathan Alexander Prestes Fruscalso, Roger Pizzato Nunes, Alexandre Bonatto, and Mirko Salomón Alva-Sánchez. 2022. "Highly Enriched Uranium-Free Medical Radioisotope Production Methods: An Integrative Review" Applied Sciences 12, no. 24: 12569. https://doi.org/10.3390/app122412569

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