Next Article in Journal
Evaluating Software Tools for Lipid Identification from Ion Mobility Spectrometry–Mass Spectrometry Lipidomics Data
Previous Article in Journal
Synthesis and Evaluation of 68Ga-Labeled (2S,4S)-4-Fluoropyrrolidine-2-Carbonitrile and (4R)-Thiazolidine-4-Carbonitrile Derivatives as Novel Fibroblast Activation Protein-Targeted PET Tracers for Cancer Imaging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Comparison of Radionuclide Impurities Activated during Irradiation of 18O-Enriched Water in Tantalum and Silver Targets during the Production of 18F in a Cyclotron

by
Teresa Jakubowska
1,2,3,
Magdalena Długosz-Lisiecka
1,* and
Michał Biegała
2,3
1
Institute of Applied Radiation Chemistry, Faculty of Chemistry, Lodz University of Technology, Wróblewskiego 15, 90-924 Łódź, Poland
2
Department of Medical Physics, Copernicus Memorial Hospital in Lodz Comprehensive Cancer Center and Traumatology, Pabianicka 62, 93-513 Łódź, Poland
3
Department of Medical Imaging Technology, Medical University of Lodz, ul. Lindleya 6, 690-131 Łódź, Poland
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(8), 3485; https://doi.org/10.3390/molecules28083485
Submission received: 3 March 2023 / Revised: 3 April 2023 / Accepted: 5 April 2023 / Published: 14 April 2023
(This article belongs to the Section Physical Chemistry)

Abstract

:
During the production of 18F, as a result of the interaction of the beam of protons and secondary neutrons with the structural elements of the target body, many radionuclide impurities are created in the cyclotron. As part of this work, we theoretically predicted which isotopes would be activated in the target tantalum or silver bodies. Subsequently, we used gamma spectrometry analysis to verify these predictions. The results were compared with the work of other authors who studied titanium and niobium as materials for making the target body. Tantalum has been evaluated as the most favorable in terms of generating radionuclide impurities during the production of 18F by irradiation of 18O-enriched water in accelerated proton cyclotrons. Only three radionuclides were identified in the tested samples: 181W, 181Hf, and 182Ta with a half-life of fewer than 120 days. The remaining reactions led to the formation of stable isotopes.

1. Introduction

The 18F isotope, most commonly used in positron emission tomography (PET), is produced in a cyclotron through bombarding the 18O-enriched water in the target casing with a beam of accelerated protons. Many factors are taken into account when designing the target body.
During the production of radionuclides, the irradiated target material in cyclotrons is most often in a liquid or gas form. During the bombardment with protons in these centers, rapid heating occurs, causing an increase in their volume and, in the case of liquid, even boiling [1,2,3,4]. In addition, the protons from the beam partially lose their energy on their way in the target’s structural elements, causing their intense heating. In practice, the working chamber of the target is a hole drilled in a cylinder made of metal. This metal must be relatively easy to machine but of a high strength and characterized by good thermal conductivity. The working chamber from the side of the beam entrance has a window made of a durable foil pressed against the working space of the target with an increased pressure, necessary to compensate for the effects of rapid heating of the irradiated medium. Increasing the strength of the target window by increasing the thickness of the foil would cause too much loss in the proton beam that passes through it. Therefore, this problem was solved by using an additional mesh supporting the thin foil. The most commonly used are Havar alloy foils with a high cobalt content supported by a mesh made of copper or aluminum [1,2,3]. Target bodies are usually made of silver, tantalum, titanium, or niobium [5,6,7,8].
The cooling of the target material and target body components is one of the main considerations when designing the target body. The target body made of silver, due to its high thermal conductivity of 415 W/(m·K), is characterized by a high heat exchange efficiency, whereas the thermal conductivity of tantalum is 57 W/(m·K), and that of titanium is only 21.9 W/(m·K). The analyzed cyclotron originally used a target body made of silver, which was then replaced with a tantalum body. The reduced heat transfer rate of tantalum was compensated for by the increased cooling of the rear face of the target body using an additional radiator. In addition, the irradiation of the water enriched in the silver body led to the formation of colloids, which resulted in a decrease in the efficiency of radionuclide production. In this case, the entire item had to be removed from the accelerator and cleaned to have its functionality restored. Making the target body composed of tantalum extended the life of the element and reduced the number of impurities in the liquid target material.
This work discusses the influence of the used target body on radionuclide contaminants in the irradiated material. Such impurities are most often formed when the beam passes through the target window, but they can also arise when protons with sufficient energy hit the walls of the target working chamber. Earlier studies analyzed the formation of such impurities in the target with a housing made of silver [9]; there is a discussion in the literature regarding isotopes activated in titanium and resulting from irradiation of the beam of protons accelerated in the cyclotron [10,11,12]. The activation processes resulting in the production of radionuclides also occur in the elements of linear medical accelerators in nuclear reactions caused by photons accelerated to high energies and secondary neutrons [13,14,15]. In the following paper, theoretically possible nuclear reactions with tantalum will be discussed, which result in the formation of stable and radioactive isotopes that are impurities during the production of isotopes for positron emission tomography. The following reactions were analyzed: (n,p), (p,2n), (p,a), (p,d), (n,p), and (n,g). Theoretical predictions will be compared with the results of gamma spectrometry measurements of the set of columns for the purification of fluorodeoxyglucose [18F]FDG.

2. Results

The results of the nuclear reactions (n,p), (p,2n), (p,a), (p,d), (n,p), and (n,g) with the isotopes from elements used in both target bodies were theoretically predicted. The tables below show the possible nuclear reactions taking place in target bodies together with the maximum value of the cross-sections of these reactions with the energy values. The analyzed energy range spans from the reaction threshold energy to the maximum proton energy of 11 MeV (Table 1 and Table 2). Natural silver contains two stable isotopes: 107Ag and 109Ag. The first isotope is 51.839%, and the second is 48.161%. Natural tantalum is 99.988% stable isotope 181Ta.
The predicted radioactive isotopes that can be formed in the natural silver target casing are 107Pd, 109Pd, 106Ag, 108Ag, 109Cd, and 107Cd (Table 1). The presence of the cadmium isotope 109Cd was confirmed in a previous study [9]. In the case of a tantalum target casing, possible radioactive impurities are isotopes 181W, 180W, and 181Hf (Table 2). The tests performed with gamma spectrometry confirmed the presence of theoretically predicted cadmium isotopes in the target made of silver in the samples from the production of FDG. Due to a very short half-life of the theoretically predicted silver isotopes, their presence could not be confirmed during measurements performed several dozen hours after the end of the irradiation. The presence of the theoretically predicted isotopes of palladium (107Pd, 109Pd) has also not been confirmed. These are isotopes decaying with the emission of the β-electron, characterized by single emission lines of gamma radiation, and their reaction cross-section is an order of magnitude lower than that of the confirmed cadmium isotopes.
In our study, the target window was made of Havar alloy foil, and many radioisotope impurities originating from nuclear reactions between protons from the beam and secondary neutrons and atoms of elements constituting the Havar alloy were identified in the tested samples (Table 3). These are 55,56,57Co, 95,95m,96Tc, 52Mn, and 182,182m,183,184,186Re. Characteristic photon peaks for these radionuclides were confirmed in the spectrometric spectra of the set of columns used in the production of the target housings made of both silver and tantalum (Figure 1 and Figure 2). The presence of remium isotopes 182Re, 182mRe, 183Re, 184Re, 184mRe, and 186Re makes identification of 182Ta difficult because most of the peaks are characteristic of tantalum coincide with those of gamma radiation characteristic of at least one isotope of remium.

3. Materials and Methods

The cyclotron used in the work is located in the Provincial Multi-Specialty Center of Oncology and Traumatology in Łódź [3,16,17]. The Eclipse™ RD cyclotron (Siemens, Warsow, Poland) used is a bare-shielded cyclotron that accelerates proton energy to 11 MeV and produces 18F and 11C positron-emitting radioisotopes. The 18F isotope is formed during the bombardment of 18O-enriched water in the 18O(p,n)18F reaction, and the target casing was initially made of silver and then replaced with another made of tantalum during the upgrade of the cyclotron [18]. Tantalum target for the production of F-18 was placed in a visible, empty working chamber and filled with oxygen-18-enriched water and irradiated during cyclotron operation (Figure 3).
Typical irradiation parameters include:
  • Enrichment of 18O, typically >95%.
  • Chemically pure enriched water 18O, above 99.99%.
  • Target volume of 2.5 mL.
  • A beam of protons with an energy of 11 MeV.
  • Beam current of 60 μA.
  • Irradiation time of 90 min.
Figure 3. The target for the production of the F-18 (RDS 111 & HP ECLIPSE Target, Simens, Warsaw, Poland) and the scheme of its construction. A—Copper Hex Grid, B—target window (Havar foil), C—Target Body (silver or tantalum), D—Target Support Cylinder Set with radiator and water cooling.
Figure 3. The target for the production of the F-18 (RDS 111 & HP ECLIPSE Target, Simens, Warsaw, Poland) and the scheme of its construction. A—Copper Hex Grid, B—target window (Havar foil), C—Target Body (silver or tantalum), D—Target Support Cylinder Set with radiator and water cooling.
Molecules 28 03485 g003
An 11 MeV proton beam accelerated in a cyclotron passes through a 50 μm thick Havar window almost completely (99.999%), losing only 3.3 keV of its energy [19]. During the collisions, some of the protons from the beam change direction and hit the target’s body, causing nuclear side reactions that produce the radionuclide contaminants analyzed in this work. Protons mainly lose their energy in enriched water. However, this energy is sufficient to produce 18F in the 18O(p, n)18F reaction, because the cross-section of this reaction is maximal at 5–6 MeV.
In theoretical considerations, the open database system of the International Atomic Energy Agency (IAEA) TENDL-2019 (TALYS Evaluated Nuclear Data Library version 2019) was used to predict possible nuclear reactions, and the Nuclear Data Section database was used to determine the stability and decay methods of the resulting isotopes (https://www-nds.iaea.org (accessed on 7 February 2023)). Kernel model is a computer code applied for predicting and analyzing nuclear reactions. Code simulates reactions involving neutrons, gamma rays, protons, deuterons, tritons, helions, and alpha particles in the energy range from 1 keV to 200 MeV [20,21,22,23,24,25].
The radionuclide identity can be confirmed by obtaining a gamma spectrum or by measuring the half-life of the product. Therefore, for the identification and quantification of radioisotopes, a high-resolution, low-background gamma spectrometry system was used, equipped with an HPGe detector with a relative efficiency of up to 30%. The analysis was carried out using a dedicated Genie 2000 software and an ISOCS/Lab SOCS supporting software.
Samples from nine different days of routine fluorodeoxyglucose production were tested using a tantalum target housing with a Havar foil window. The tested samples were elements of the filtering system applied for the synthesis of [18F]FDG: a separation column used to capture fluorine ions from the water after irradiation—QMA (Quaternary ammonium anion exchange) and a set of filtration columns used to purify the finished FDG. The filter set consisted of five columns used in the on-chip set: IC-H (cation exchange resin), two IC-HCO3 (anion exchange resin), AL-N, and C18. Each of the column sets was used in only one production run. They were subjected to a spectrometric analysis 30–48 h after the end of the synthesis. The results were compared with the results of analogous five sets obtained in the production of FDG with the use of silver-body targets.

4. Discussion

In this work, a gamma radiation spectrometer was used for a quantitative and qualitative analysis of the elements of the FDG synthesis modules, such as a separation column and a set of filtration columns. Similar analyses were carried out for the target body made of silver with windows made of Havar foil, and the results of these analyses were described in the works of Ferguson (2011) [13], Bowden (2009) [14], and Marengo (2008) [20]. Mochizuki (2006) [11] also analyzed the elements of the synthesis modules, but for the target casing made of titanium, similarly to the team of Guarino et al. (2007) [12]. Meanwhile, Köhler (2013) [15] analyzed samples in the form of water enriched in 18O after irradiation in the production of FDG in a target with a window from niobium. Researchers using silver target casings identified the isotopes 109Cd, 105Ag, 106mAg, and 110mAg as radionuclide impurities, the source of which were reactions occurring in the immediate vicinity of the target (Table 4).
Among the isotopes indicated in this work, the presence of 109Cd and 107Cd detected in our previous studies was confirmed [3]. Two isotopes, 46Sc and 48V, were identified during the period of use of the titanium case. The sources of these radionuclide impurities are the reactions 46Ti(n,p)46Sc, 48Ti(p,n)48V, and 48Ti(p,2n)48V. Kambali et al. (2020) [10], in their work, predicted the possibility of generating additional 10 radioisotopes when using a titanium target casing: 43,44,47Sc,45,46,47,49,50V and 45,51Ti. Köhler, using a niobium target window, additionally identified isotopes 89Zr, 92mNb, and 93mMo. In this work, we are the first to confirm the presence of the indicated isotope impurities in practice during the production of FDG in the tantalum target casing as a source of radionuclide impurities and to quantitatively confirm their contribution. We had a unique opportunity to compare the effect of changing the target on the formation of impurities in practice, when after seven years of use, the silver target was replaced with an identical tantalum one, which has already been used for two years. Spectrometric analysis identified three new radionuclide impurities: 181W, 181Hf, and 182Ta. In the previous target, the emerging cadmium isotopes were the dominant impurity, which is very clearly visible in the presented picture (Figure 2). In the current system, the most dominant gamma-ray peak comes from 182Ta, while 181W and 181Hf are difficult to identify. The nuclear reaction cross-sections for the production of these two isotopes are incomparable to the cross-sections for the production of 182Ta.
The main source of radionuclide impurities arising during the production of 18F is the window closing the target working chamber. The type of formed radionuclides depends on the used material; the most common are Havar foils, but also niobium and titanium are used. The energy of the proton beam also affects the number of pollutants formed.

5. Conclusions

The sources of radiation in the region of cyclotron are the basic nuclear reactions (p,n) and (d,n). Examples of these reactions are 18O(p,n)18F, 15N(p,n)15O, 12C(d,n)13N, and 14N(d,n)15O. Neutron radiation is also observed during the production of the 11C isotope in the 14N(p, α)11C reaction in which no neutrons are produced. This is due to the fact that, in addition to the basic reactions, neutrons can be generated by other secondary reactions caused by the interaction with the elements of the cyclotron directly in the path of the beam of accelerated charged particles.
Among the analyzed metals, tantalum is the optimal material in terms of the production of radionuclide impurities during the production of 18F by the irradiation of 18O-enriched water with protons in accelerated cyclotrons. Only three radionuclides with half-lives of fewer than 120 days were identified in the tested samples. In the case of silver body target impurity, the half-life of the identified 109Cd is 462 days. Therefore, a new cyclotron body target based on tantalum has been evaluated, as it promises a new material potentially classified as intermediate or low-level radioactive waste in future.

Author Contributions

Conceptualization, T.J.; Methodology, T.J.; Software, M.D.-L.; Validation, M.D.-L.; Formal analysis, T.J.; Resources, M.B.; Data curation, M.D.-L.; Writing—original draft, T.J.; Writing—review & editing, M.B.; Supervision, M.D.-L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. O’Donnell, R.G.; Vintro, L.L.; Duffy, G.J.; Mitchell, P.I. Measurement of the residual radioactivity induced in the front foil of a target assembly in a modern medical cyclotron. Appl. Radiat. Isot. 2004, 60, 539–542. [Google Scholar] [CrossRef] [PubMed]
  2. Martinez-Serrano, J.J.; Diez de Los Rios, A. Predicting induced activity in the Havar foils of the (18)F production targets of a PET cyclotron and derived radiological risk. Health Phys. 2014, 107, 103–110. [Google Scholar] [CrossRef]
  3. Długosz-Lisiecka, M.; Jakubowska, T.; Zawada, A. High-Level Radioactive Wastes from 18F and 11C Isotopes Production. J. Hazard. Toxic Radioact. Waste 2021, 25, 04020072. [Google Scholar] [CrossRef]
  4. Berridge, M.S.; Kjellström, R. Designs and use of silver [18O]water targets for [18F]fluoride production. Appl. Radiat. Isot. 1999, 50, 699–705. [Google Scholar] [CrossRef]
  5. Jacobson, O.; Kiesewetter, D.O.; Chen, X. Fluorine-18 radiochemistry, labeling strategies and synthetic routes. Bioconjug. Chem. 2015, 26, 1–18. [Google Scholar] [CrossRef] [Green Version]
  6. Helmeke, H.-J.; Harms, T.; Knapp, W.H. A water target with beam sweep for routine fluorine-18 production. Appl. Radiat. Isot. 2001, 54, 753–759. [Google Scholar] [CrossRef]
  7. Zeisler, S.K.; Becker, D.W.; Pavan, R.A.; Moschel, R.; Rühle, H. A water-cooled spherical niobium target for the production of [18F]fluoride. Appl. Radiat. Isot. 2000, 53, 449–453. [Google Scholar] [CrossRef]
  8. Siikanen, J.; Ohlsson, T.; Medema, J.; Van-Essen, J.; Sandell, A. A niobium water target for routine production of [18F]Fluoride with a MC 17 cyclotron. Appl. Radiat. Isot. 2013, 72, 133–136. [Google Scholar] [CrossRef]
  9. Długosz-Lisiecka, M.; Jakubowska, T.; Zbrojewska, M. Formation of 107Cd radionuclide impurities during 18F production. Appl. Radiat. Isot. 2019, 147, 48–53. [Google Scholar] [CrossRef]
  10. Kambali, I. Comprehensive theoretical studies on stable and radioactive isotopes produced by proton irradiation of titanium target. J. Phys. Conf. Ser. 2020, 1572, 012013. [Google Scholar] [CrossRef]
  11. Mochizuki, S.; Ogata, Y.; Hatano, K.; Abe, J.; Ito, K.; Ito, Y.; Nishino, M.; Miyahara, H.; Ishigure, N. Measurement of the Induced Radionuclides in Production of Radiopharmaceuticals for Positron Emission Tomography (PET). J. Nucl. Sci. Technol. 2006, 43, 348–353. [Google Scholar] [CrossRef]
  12. Guarino, P.; Rizzo, S.; Tomarchio, E.; Greco, D. Gamma-ray spectrometric characterization of waste activated target components in a PET cyclotron, Cyclotrons and their applications. In Proceedings of the 18th International Conference, Cyclotrons 2007, Giardini Naxos, Italy, 1–5 October 2007. [Google Scholar]
  13. Ferguson, D.; Orr, P.; Gillanders, J.; Corrigan, G.; Marshall, C. Measurement of long lived radioactive impurities retained in the disposable cassettes on the Tracerlab MX system during the production of [18F]FDG. Appl. Radiat. Isot. 2011, 69, 1479–1485. [Google Scholar] [CrossRef]
  14. Bowden, L.; Vintró, L.; Mitchell, P.J.; O’Donnell, R.G.; Seymour, A.M.; Duffy, G.J. Radionuclide impurities in proton-irradiated [18O]H2O for the production of 18F−: Activities and distribution in the [18F]FDG synthesis process. Appl. Radiat. Isot. 2009, 67, 248–255. [Google Scholar] [CrossRef]
  15. Köhler, M.; Degering, D.; Zessin, J.; Füchtner, F.; Konheiser, J. Radionuclide impurities in [18F]F− and [18F]FDG for positron emission tomography. Appl. Radiat. Isot. 2013, 81, 268–271. [Google Scholar] [CrossRef]
  16. Długosz-Lisiecka, M.; Biegała, M.; Jakubowska, T. Activation of medical accelerator components and radioactive waste classification based on low beam energy model Clinac 2300. Radiat. Phys. Chem. 2023, 205, 110730. [Google Scholar] [CrossRef]
  17. Jakubowska, T.; Długosz-Lisiecka, M. Estimation of effective dose using two ICRP criteria, applied to radiation protection of personnel in an unshielded PET cyclotron facility. Radiat. Phys. Chem. 2020, 171, 108688. [Google Scholar] [CrossRef]
  18. Francis, F.; Wuest, F. Advances in [18F]Trifluoromethylation Chemistry for PET Imaging. Molecules 2021, 26, 6478. [Google Scholar] [CrossRef]
  19. Sublet, J.-C.; Dzysiuk, N.; Fleming, M.; van der Marck, S. TENDL: Complete Nuclear Data Library for Innovative Nuclear Science and Technology. Nucl. Data Sheets 2019, 155, 1–55. [Google Scholar] [CrossRef]
  20. Marengo, M.; Lodi, F.; Magi, S.; Cicoria, G.; Pancaldi, D.; Boschi, S. Assessment of radionuclid icimpurities in2-[18F]fluoro-2-deoxy-d-glucose ([18F]FDG) routine production. Appl. Radiat. Isot. 2008, 66, 295–302. [Google Scholar] [CrossRef]
  21. Satyamurthy, N.; Amarasekera, B. Phelps Tantalum [18O]Water Target for the Production of [18F]Fluoride with High Reactivity for the Preparation of 2-Deoxy-2-[18F]Fluoro-D-Glucose. Mol. Imaging Biol. 2002, 4, 65–70. [Google Scholar] [CrossRef]
  22. Kambali, I.; Suryanto, H. Recoiled and sputtered radioactive impurities in 11 MeV proton-based F-18 production; recoiled and sputtered radioactive impurities in 11 MeV proton-based F-18 production. Int. J. Technol. 2019, 10, 300. [Google Scholar] [CrossRef] [Green Version]
  23. Üncü, Y.A.; Özdoğan, H.; Şekerci, M.; Kaplan, A. Investigation of the production routes of Palladium-103 and Iodine-125 radioisotopes. Radiat. Phys. Chem. 2023, 204, 110658. [Google Scholar] [CrossRef]
  24. Deilami-Nezhad, L.; Moghaddam-Banaem, L.; Sadeghi, M.; Asgari, M. Production and purification of Scandium-47: A potential radioisotope for cancer theranostics. Appl. Radiat. Isot. 2016, 118, 124–130. [Google Scholar] [CrossRef] [PubMed]
  25. Yiğit, M.; Bostan, S.N. Study on cross section calculations for (n,p) nuclear reactions of cadmium isotopes. Appl. Radiat. Isot. 2019, 154, 108868. [Google Scholar] [CrossRef]
Figure 1. Gamma ray spectrum of the QMA column with visible gamma peaks characteristic of Cd-107, Cd-109 isotopes from FDG production using a silver target.
Figure 1. Gamma ray spectrum of the QMA column with visible gamma peaks characteristic of Cd-107, Cd-109 isotopes from FDG production using a silver target.
Molecules 28 03485 g001
Figure 2. Gamma ray spectrum of the QMA column with visible gamma peaks characteristic of the isotopes W-181, W-180, and Hf-181 from FDG production using a tantalum target.
Figure 2. Gamma ray spectrum of the QMA column with visible gamma peaks characteristic of the isotopes W-181, W-180, and Hf-181 from FDG production using a tantalum target.
Molecules 28 03485 g002
Table 1. Theoretical analysis of nuclear reactions with silver isotopes 107Ag and 109Ag.
Table 1. Theoretical analysis of nuclear reactions with silver isotopes 107Ag and 109Ag.
Nuclear ReactionThreshold Energy [MeV]Maximum Cross-Section [mbarn]/Energy [MeV]Decay ModeHalf-Life
107Ag(p,n)107Cd2.219519.354/11EC, β+6.5 h
107Ag(p,2n)106Cd10.2295.313/11Stable-
107Ag(p,a)104Pa01.792/11Stable-
107Ag(p,d)106Ag7.3856,952/11EC, β+23.96 min
107Ag(n,g)108Ag0852/thermal neutronsEC, β+2.382 min
107Ag(n,p)107Pd024.43/11β−6.5 × 106 y
109Ag(p,n)109Cd1.0071394/9EC, β+462 d
109Ag(p,2n)108Cd8.398428/11Stable-
109Ag(p,a)106Pa0394/11Stable-
109Ag(p,d)108Ag7.0240.001/11EC, β+2.38 min
109Ag(n,g)110Ag01,784,480/thermal neutronsStable-
109Ag(n,p)109Pd0.33412.61/11β−13.59 h
Table 2. Theoretical analysis of nuclear reactions with tantalum isotope 181Ta.
Table 2. Theoretical analysis of nuclear reactions with tantalum isotope 181Ta.
Nuclear ReactionThreshold Energy Maximum Cross-Section [mbarn]/Energy [MeV]Decay ModeHalf-Life
181Ta(p,n)181W0.97664.994/9EC. β+121.2 d
181Ta (p,2n)180W7.699398.1/11Stable-
181Ta (p,a)178Hf8.860.0352/11Stable-
181Ta (p,d)180Ta5.380.000009/11Stable-
181Ta (n,g)182Ta01,784,480/thermal neutronsEC. β+114.74 d
181Ta (n,p)181Hf00.493/thermal neutronsβ−42.39 d
Table 3. Isotopes identified in the spectra with their characteristic peaks.
Table 3. Isotopes identified in the spectra with their characteristic peaks.
IsotopeHalf-LifeGamma Radiation/Intensity
52Mn5.591 d1434.09 (100%), 935.54 (94.5%), 744.23 (90%)
55Co17.53 h931.3 (75%), 477.2 (20.2%), 1408.5 (16.9%)
56Co77.23 d846.77 (99.9%), 1238.29 (66.46%), 2598.50 (16.97%), 1771.35 (15.41%), 1037.84 (14.05%)
57Co271.74 d122.06 (85.60%), 136.47 (10.68%)
95Tc20 h765.79 (93.82%)
95mTc61 d204.117 (63.20%), 582.08 (29.96%),
835.149 (26.63%)
96Tc4.28 d778.224 (99.76%), 849.93 (98%),
812.58 (82%), 1126.85 (15.2%)
107Cd6.5 h93.12 (4.8%), 828.93 (0.17%)
109Cd462.6 d88.03 (3.61%)
181W121.2 d152.32 (0.08%), 136.28 (0.03%)
181Hf43.39482.18 (80.5%), 133.02 (43.3%), 345.92 (15.12%)
182Ta114.43 d67.749 (42.9%), 1121.3 (35.24%), 1221.4 (27.23 %), 1189.04 (16.49%)
182Re64 h229.32 (25.8%), 67.85 (22.2%), 1121.3 (22.1%), 1221.1 (17.5%), 100.1 (16.5%),
1231.01 (14.9%), 169.15 (11.4 %)
182mRe14.14 h67.75 (38.3%), 1121.4 (32%), 1221.5 (25%),
1189.2 (15.1%), 100.2 (14.4%)
183Re70 d162.33 (25.1%), 46.48 (8.0%)
184Re38 d903.28 (38.1%), 792.06 (37.7%), 111.21 (17.2%), 894.76 (15.7%)
184mRe169 d104.73 (13.4%)
186Re3.718 d137.15 (9.47%)
Table 4. List of radioactive impurities identified in target bodies by various research groups.
Table 4. List of radioactive impurities identified in target bodies by various research groups.
AutorPoint of AnalysisTarget Foil/Target BodyIdentified Importunities from FoilIdentified Importunities from Target Body
Ferguson [13]Synthesis cassettesHavar/silver51Cr, 52,54Mn, 56,57,58Co, 95m,96Tc, 182,183Re109Cd
Bowden [14]Silver target body “in situ” in cyclotron, synthesis cassettes, sample of irradiated [18O]H2O, before and passing through the QMA columnHavar/silver51Cr, 52Mn, 56,57,58Co, 95m,96Tc, 183,184Re110mAg
109Cd
Mochizuki [11]Target foil, separation and filtration column Havar/titan48V, 51Cr, 52,54Mn, 56,57,58,60Co, 95m,96Tc, 183,184Re46Sc, 48V
Guarino [12]Target window foil (havar), titanium vacuum windowHavar and titanium/-H: 51Cr, 52,54Mn, 56,57,58Co,
Ti: 46Sc, 48V
Not analyzed
Köhler [15]Sample of irradiated [18O]H2ONiob/-48V, 51Cr, 52,54Mn, 55,56,57,58Co, 57Ni, 89Zr, 92mNb, 93mMo, 95,96Tc,Not analyzed
Marengo [20]Synthesis cassettes, sample of collected [18O]H2O, purification filtersHavar/silver48V, 51Cr, 52,54,56Mn, 55,56,57,58,60Co, 57Ni, 95,95m,96,98Tc, 182,182m,183,184,186Re105Ag, 106mAg, 109Cd
Długosz-Lisiecka [9]Synthesis cassettes,
QMA separation filter, purification filters
Havar/silverNot analyzed107Cd, 109Cd
in this workQMA separation filter, purification filtersHavar/silver
Havar/tantal
52Mn, 55,56,57Co, 95,95m,96Tc, 182,182m,183,184,184m,186ReAg: 107Cd, 109Cd
Ta: 181W, 181Hf, 182Ta
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

Jakubowska, T.; Długosz-Lisiecka, M.; Biegała, M. Comparison of Radionuclide Impurities Activated during Irradiation of 18O-Enriched Water in Tantalum and Silver Targets during the Production of 18F in a Cyclotron. Molecules 2023, 28, 3485. https://doi.org/10.3390/molecules28083485

AMA Style

Jakubowska T, Długosz-Lisiecka M, Biegała M. Comparison of Radionuclide Impurities Activated during Irradiation of 18O-Enriched Water in Tantalum and Silver Targets during the Production of 18F in a Cyclotron. Molecules. 2023; 28(8):3485. https://doi.org/10.3390/molecules28083485

Chicago/Turabian Style

Jakubowska, Teresa, Magdalena Długosz-Lisiecka, and Michał Biegała. 2023. "Comparison of Radionuclide Impurities Activated during Irradiation of 18O-Enriched Water in Tantalum and Silver Targets during the Production of 18F in a Cyclotron" Molecules 28, no. 8: 3485. https://doi.org/10.3390/molecules28083485

Article Metrics

Back to TopTop