logo

Progress of photonuclear cross-sections for medical radioisotope production at the SLEGS energy domain

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Progress of photonuclear cross-sections for medical radioisotope production at the SLEGS energy domain

Xuan Pang
Bao‑Hua Sun
Li‑Hua Zhu
Guang‑Hong Lu
Hong‑Bo Zhou
Dong Yang
Nuclear Science and TechniquesVol.34, No.12Article number 187Published in print Dec 2023Available online 30 Nov 2023
52308

Photonuclear reactions using a laser Compton scattering (LCS) gamma source provide a new method for producing radioisotopes for medical applications. Compared with the conventional method, this method has the advantages of a high specific activity and less heat. Initiated by the Shanghai Laser Electron Gamma Source (SLEGS), we conducted a survey of potential photonuclear reactions, γ, n, (γ, p), and (γ, γ′) whose cross-sections can be measured at SLEGS by summarizing the experimental progress. In general, the data are rare and occasionally inconsistent. Therefore, theoretical calculations are often used to evaluate the production of medical radioisotopes. Subsequently, we verified the model uncertainties of the widely used reaction code TALYS-1.96, using the experimental data of the 100Moγ, n99Mo, 65Cuγ, n64Cu, and 68Zn(γ, p)67Cu reactions.

Medical radioisotopePhotonuclear reactionLCSCross section
1

Introduction

Radioisotopes are widely used for the diagnosis and therapy of different diseases owing to their nuclear-physical properties [1, 2]. Diagnostic radioisotopes can provide functional and metabolic information for the early treatment of diseased regions that have not yet undergone morphological or structural changes. Currently, positron emission tomography (PET) [3] and single-photon emission computed tomography (SPECT) [4] are the two major diagnostic techniques. Short-lived β+-emitting radioisotopes are often used for PET. Typical examples include 11C, 13N, 15O, and 18F, which have half-lives of 20, 10, 2, and 110 min, respectively. For SPECT, γ-ray emitting radioisotopes are frequently used, of which 99mTc with a half-life of 6 h [5] is the most common radioisotope tracer. Therapeutic radioisotopes can be combined with targeted drugs to achieve the precise removal of small diseased regions without excessive doses to normal tissues. Therefore, radioisotopes that emit low-range highly ionizing radiation are of significant interest. The β--particle emitting radioisotopes (e.g. 32P, 89Sr, 90Y, 131I, 177Lu, and 188Re), Auger electron cascades (e.g. 103Pd and 125I) and α-particle emitting radioisotopes (e.g. 211At, 212Bi, 223Ra, 225Ac, and 227Th), with a high linear energy transfer in tissue, are suitable for therapy.

More than 40 million nuclear medicine procedures are performed annually, and the demand for radioisotopes is increasing by 5% annually [6]. Currently, the production of medical radioisotopes relies primarily on nuclear reactors and cyclotrons. Thermal neutron-induced fission produces neutron-rich radioisotopes in nuclear reactors. The advantage of this method is the possibility of producing high levels of total and specific activities. However, the production of desired radioisotopes is accompanied by a considerable amount of long-lived radioactive waste, which raises numerous safety and security concerns. In addition, many nuclear reactors used for medical radioisotope production are more than 50 years old and may be shut down in the near future [7]. For example, a vast majority of reactors producing 99Mo are expected to shut down by 2030. Compared with reactors, cyclotrons typically produce neutron-deficient radioisotopes by charged particle reactions accompanied by less radioactive waste. In addition, they have the advantage of being more compact, allowing their placement near hospitals, and making them a useful tool for producing radioisotopes with short half-lives that range from several minutes to hours.

Another promising method for radioisotope production involves photonuclear reactions, mainly including γ, n, (γ, p), and (γ,γ’). This is considered an alternative to radioisotope production in reactors and cyclotrons or the only method of production for certain radioisotopes. γ, n reactions can produce β+-emitting radioisotopes for PET imaging, such as 11C, 13N, 15O, and 18F. These radioisotopes are produced by cyclotron-based (p,n) or (p, γ) reactions. (γ, p) reactions are suitable for producing β--emitting radioisotopes that are currently mostly produced in reactors. High-intensity γ beams are required to obtain adequate yield using this method. With the development of laser Compton-scattering (LCS) gamma sources, the production of radioisotopes via photonuclear reactions has attracted considerable attention [8-17].

In contrast to conventional bremsstrahlung gamma sources based on electron linear accelerators, LCS gamma sources have the advantages of a high photon flux and excellent monochromaticity. The advantages of photonuclear reactions using LCS gamma sources for radioisotope production are as follows:

• Differing from bremsstrahlung gamma sources, the LCS gamma sources significantly reduce the heat per useful reaction rate as its γ rays in the energy range of interest are not accompanied by an intense low-energy tail. Moreover, the LCS gamma sources have the ability to selectively tune photons to energies of interest. A high specific activity can be achieved by matching the photon energy to the giant dipole resonance (GDR) peak.

• Differing from the charged-particle induced reactions, the photonuclear reactions generate less heat as their energy deposition in targets via gamma-matter interactions is significantly smaller. Thus, the target can be thicker and requires considerably less cooling [9]. Moreover, the photonuclear reactions have the capability to simultaneously irradiate multiple targets. In this manner, it can maximize the utilization of γ beams and produce multiple radioisotopes simultaneously [8].

The Shanghai Laser Electron Gamma Source (SLEGS) employs the LCS technique. It can generate γ beams ranging from 0.25–21.7 MeV with a full-spectrum flux of 105–107 s-1 [18] and the best possible bandwidth of γ beams of 5%–15% after passing through a dual collimation system [19]. Using collimation technology, the size of an γ beam can be continuously adjusted to within Φ25 mm [20]. The monochromaticity and high intensity of γ beams combined with detector spectrometers can be used to measure the photonuclear reaction cross-section [21-24]. One of the main topics of SLEGS is the measurement of the key photonuclear reaction cross-section relevant for medical radioisotope investigations, providing crucial nuclear data for the photonuclear method.

In this study, we aim to estimate the production of medical radioisotopes in the energy range of 0.25–21.7 MeV in the SLEGS domain. The remainder of this paper is organized as follows. In Sect. 2, we summarize the radioisotopes that can be produced via (γ, n), (γ, p), and (γ,γ’) reactions and discuss their experimental feasibility. Owing to the scarcity of cross-sectional data regarding photonuclear reactions, theoretical calculations are often used to assess the yields of medical radioisotopes. In Sect. 3, the cross-sectional data of the 100Mo(γ, n), 65Cu(γ, n), and 68Zn(γ, p) reactions in the SLEGS energy region were investigated to produce 99Mo, 64Cu, and 67Cu radioisotopes. Finally, we provide a summary in Sect. 4.

2

Medical radioisotopes produced in photonuclear reactions

In this section, we discuss potential γ, n, (γ, p), and (γ, γ’) reactions for the production of medical radioisotopes. Crucial considerations include the natural abundance of the target isotopes, half-life of the radioisotopes, and available experimental data. Generally, it is preferable to have a target isotope with a dominant abundance (e.g., greater than 10%), and a suitable half-life (several tens of minutes to days) for the produced radioisotope. However, the lack of experimental data may limit the medical applications of radioisotopes.

Different methods are available for measuring photonuclear reactions, including in-beam and offline measurements. Four π 3He neutron detection arrays, charged particle detectors, and nuclear resonance fluorescence spectrometers are commonly used for in-beam measurements of γ, n, (γ, p), and (γ, γ’) reactions. Offline measurement is suitable for determining all three types of photonuclear reactions only in cases where the residual decay has an unambiguous γ-ray and the lifetime of the residual nucleus is within a range of minutes to hours.

2.1
(γ, n) reactions

Radioisotopes can be most efficiently produced by exciting the target nuclide into a GDR using photons. The GDR is characterized by a large peak cross-section and broad width, resulting in a large integral cross-section. The shape of the GDR follows a Lorentzian distribution with a spreading width of approximately 5 MeV [25-27]. Generally, a considerable portion of the total photonuclear cross-section is known to originate from the γ, n reactions, which increases as the charge number of the target nuclide increases, making the following chemical separation easier. For example, the 100Moγ, n reaction accounts for more than 70% of the total photonuclear reactions within the energy range of 8.4–15 MeV. Therefore, medium- and heavy-mass nuclei are ideal for production via γ, n reactions [28].

Table 1 summarizes the 15 medical radioisotopes produced by (γ, n) reactions [8, 14, 29-31]. Excluding 100Mo, the natural abundance of these reaction targets exceeds 10%. A higher natural abundance results in a higher yield and lower impurity content. The seventh column indicates the current status of the EXFOR database [32]. The experimental data for only nine radioisotopes are presented in the Appendix. Among these, the 19Fγ, n reaction has only one data set that was measured in the 1960s. The data for the 90Zrγ, n, 100Moγ, n, and 187Reγ, n reactions do not cover the entire GDR energy region. There is a significant discrepancy between the two sets of data for the 65Cuγ, n reaction: It is possible to determine the relevant cross-sections at SLEGS.

Table 1
Medical radioisotope AX production via the A+1X(γ,n)AX reactions [8, 14, 29-31].
A+1Y Nat. abu. (%) AX T1/2 Decay mode Eth (MeV) EXFOR data Medical application Exp. method γ-ray (keV)
12C 98.93 11C 20.364 m %ε=100 18.7 Yes PET a;b 511
14N 99.63 13N 9.965 m %ε=100 10.5 Yes PET a;b 511
16O 99.757 15O 122.24 s %ε=100 15.6 Yes PET a -
19F 100 18F 109.77 m %ε=100 10.4 Yes PET a;b 511
45Sc 100 44Sc 3.97 h %ε=100 11.3 No PET a;b 511;1157
63Cu 69.15 62Cu 9.673 m %ε=100 10.8 Yes PET a;b 511
65Cu 30.85 64Cu 12.7 h %ε=61.5 9.91 Yes PET;Radiotherapy a;b 511
        -=38.5
69Ga 60.108 68Ga 67.71 m %ε=100 10.3 No PET a;b 511;1077
90Zr 51.45 89Zr 78.41 h %ε=100 11.9 Yes PET a -
100Mo 9.82 99Mo 65.976 h -=100 8.29 Yes SPECT a;b 740;778
104Pd 11.14 103Pd 17.0 d %ε=100 10.0 No Radiotherapy a -
166Er 33.503 165Er 10.36 h %ε=100 8.47 No Radiotherapy a -
170Er 14.91 169Er 9.4 d -=100 7.25 No Radiotherapy a -
187Re 62.60 186Re 3.7 d -=92.53 7.36 Yes SPECT;Radiotherapy a -
        %ε=7.47
193Ir 62.70 192Ir 73.8 d -=95.24 7.77 No Radiotherapy a -
        %ε=4.76          
Show more
The natural abundance of A+1Y is provided. The produced medical radioisotopes AX are characterized by the half-lives T1/2 and decay modes. Here, %ε represents the probabilities of nuclear decay via electron capture or β+ decay. %β- represents the probabilities of β- decay. Eth represents the threshold energy of the reaction. The seventh column indicates whether the A+1X(γ,n)AX reactions have experimental data in the EXFOR database [32]. a and b indicate that the radioisotope can be studied by in-beam and off-line measurements, respectively. The last column presents the probable γ-rays of interest in the offline measurement. The data are obtained from Ref. [33].
2.2
(γ, p) reactions

Despite the nucleus being excited beyond the proton separation energy of the photons, it does not necessarily lose a proton. Only for excitations well beyond the proton separation energy, the proton can acquire sufficient kinetic energy to effectively cross the Coulomb barrier. However, in these cases, the excitation energy typically exceeds the separation energies of one or two neutrons. Therefore, the neutron emission channel competes with the proton emission channel. For light and certain medium-mass nuclei, the cross-sections of the (γ, p) reactions are comparable to and occasionally exceed those of the γ, n reactions, owing to the shell structure of the nuclei. Certain heavier β- emitters for radionuclide therapy can also be produced by (γ, p) reactions, but the increasing Coulomb barrier leads to the production of small cross-sections. (γ, p) reactions result in the daughter and parent isotopes being chemically different. Several advanced chemical-separation techniques have been developed for this purpose. For example, a simple and reproducible three-ion exchange matrix approach was developed to separate 67Cu from 68Zn [34]. Separation techniques exist for other pairs, such as Ti/Sc, Zr/Y, and Hf/Lu [35-37].

The potential (γ, p) reactions used for nuclear medicine are listed in Table 2. However, the experimental data regarding this topic are scarce. Only limited data for 43K, 67Cu, and 177Lu has been obtained by using the bremsstrahlung gamma source, as detailed in the Appendix. No error bars are shown for the 43K data, whereas relatively large errors ranging from typically 10% to 100% for 67Cu and 177Lu have been demonstrated.

Table 2
Same as Table 1 but for medical radioisotopes AX production via the A+1Y(γ,p)AX reactions [8, 38-40].
A+1Y Nat. abu. (%) AX T1/2 Decay mode Eth (MeV) EXFOR data Medical application Exp. method γ-ray (keV)
44Ca 2.09 43K 22.3 h -=100 12.1 Yes Radiotherapy a;b 373;617
48Ti 73.72 47Sc 3.35 d -=100 11.4 No Radiotherapy a -
58Ni 68.077 57Co 271.74 d %ε=100 8.17 No Radiotherapy a -
68Zn 18.45 67Cu 61.83 h -=100 9.97 Yes Radiotherapy a;b 861;2502
91Zr 11.22 90Y 64.053 h -=100 8.68 No Radiotherapy a -
106Pd 27.33 105Rh 35.36 h -=100 9.34 No Radiotherapy a;b 319
112Sn 0.97 111In 2.805 d %ε=100 7.55 No SPECT;Radiotherapy a -
132Xe 26.9 131I 8.025 d -=100 9.12 No Radiotherapy a -
162Dy 25.475 161Tb 6.89 d -=100 8.00 No SPECT;Radiotherapy a -
167Er 22.869 166Ho 26.824 h -=100 7.50 No SPECT;Radiotherapy a -
178Hf 27.28 177Lu 6.647 d -=100 7.34 Yes SPECT;Radiotherapy a;b 208
189Os 16.15 188Re 17 h -=100 7.25 No SPECT;Radiotherapy a -
Show more
Nat. abu. indicate the natural abundance of A+1Y. The seventh column indicates whether experimental data regarding the A+1Y(γ,p)AX reactions is available in the EXFOR database [32].
2.3
(γ, γ’) reactions

A nuclear isomer is the metastable state of an atomic nucleus, in which one or more nucleons (protons or neutrons) occupy higher energy levels than the ground state of the same nucleus. The lifetime and excitation energy are two important properties of nuclear isomers. The excitation energy between the isomer and ground state is characteristic and can be used to identify nuclides. Nuclear isomers can be deexcited by the emission of γ rays and/or the conversion of electrons to the ground state. The γ-emitting isomers can be used as radioactive labels for SPECT imaging. Nuclear isomers emitting low-energy Auger electrons are potential radioisotopes for targeted therapy.

Conventional production methods, such as (n, γ) reactions, have relatively low yields because the dominant part of the production proceeds directly to the nuclear ground state with a spin closer to that of the target isotope. Using monochromatic and small-bandwidth LCS photons, transitions from a stable or long-lived nuclear ground state to higher energy levels can be selectively excited. Such levels serve as gateway states, which then partially decay to the isomeric state directly or via a cascade. The (γ, γ′) reaction is equivalent to storing the energy of an incident photon in an isotope that acts as a container.

Table 3 lists the six isomers that can be produced by (γ, γ′) reactions for medical applications. Excluding 117mSn and 176mLu, the experimental data for the other radioisotopes are detailed in the Appendix. The dataset of 103mRh did not have error bars, and two different results were reported from 6 to 23 MeV. For 113mIn, two cross-sectional datasets differ by more than 300 times at 8 MeV. For 115mIn, there were significant discrepancies between multiple datasets. The data of 195mPt have a large relative error as high as 100% at 3.5, 7.5, and 8 MeV.

Table 3
Same as Table 1 but for medical isomers AmX production via the AX(γ,γ)AmX reactions [8, 13].
AX Nat. abu. (%) AmX T1/2m Em/Eth (keV) Decay mode EXFOR data Medical application
103Rh 100 103mRh 56.114 m 39.8 %IT=100 Yes Radiotherapy
113In 4.29 113mIn 99.476 m 391.7 %IT=100 Yes SPECT;Radiotherapy
115In 95.71 115mIn 4.486 h 336.2 %IT=95 Yes SPECT;Radiotherapy
          -=5
117Sn 7.68 117mSn 13.76 d 314.6 %IT=100 No Radiotherapy
176Lu 2.599 176mLu 3.664 h 122.8 -=99.90 No Radiotherapy
          %ε=0.09
195Pt 33.78 195mPt 4.01 d 259.3 %IT=100 Yes Radiotherapy
Show more
The isomeric states AmX are characterized by the energy Em, half-lives T1/2m, and decay modes. Here, %IT represents the probabilities of nuclear decay via isomeric transitions. Eth equals Em. The seventh column indicates whether the AX(γ,γ)AmX reactions have experimental data in the EXFOR database [32]. AX(γ,γ)AmX reactions can be investigated by off-line measurements where the γ-ray energy of interest is equal to Em.
3

Evaluation of cross-sections for 99Mo, 64Cu, and 67Cu production

The cross-section data were used to calculate the expected yield of the radioisotopes for a given thickness and enrichment of the target material, and to determine the optimum energy range for the production of the desired radioisotope and the level of radioisotopic impurities. However, the slow development of gamma sources and small cross-sections of photonuclear reactions resulted in a relatively small number of experimental studies regarding this cross-sectional data. There is often scattering between different experimental datasets. As a result, the cross-sections predicted by the theoretical models play a key role in the production of medical radioisotopes.

The TALYS code [41] offers a unified approach for calculating nuclear reactions that involve neutrons, photons, protons, deuterons, tritons, 3He, and α particles in the keV-200 MeV energy range and for target nuclides with masses of 5 and 339. The code outputs a set of reaction data, for example cross-sections, energy spectra, and angular distributions of the emitted particles, etc.. Numerous studies have tested the TALYS code and demonstrated that it has reliable predictive power in terms of the nuclear reaction calculations [42-44]. In the TALYS code, the decay of the compound nuclear state from the photonuclear reaction was treated using the Hauser-Fenshbach (HF) statistical model [45]. The nuclear-level density and γ strength function were the main input parameters. In this study, we adapted the TALYS-1.96 code. It implemented six different nuclear level density (NLD) models and nine different models of γ strength functions (γSF), as shown in Table 4. The NLD and γSF models employ phenomenological and microscopic approaches, respectively.

Table 4
The descriptions of NLD and γSF models in the TALYS-1.96 code
Input parameter Detailed description
ldmodel 1 (default) Constant Temperature + Fermi gas model
ldmodel 2 Back-shifted Fermi gas model
ldmodel 3 Generalised Superfluid model
ldmodel 4 Skyrme-Hartree-Fock-Bogoluybov level densities from numerical tables
ldmodel 5 Gogny-Hartree-Fock-Bogoluybov (GHFB) level densities from numerical tables
ldmodel 6 Temperature-dependent GHFB level densities from numerical tables
Strength 1 (default for incident neutrons) Kopecky-Uhl generalized Lorentzian
Strength 2 (default for other incident particles) Brink-Axel Lorentzian
Strength 3 Hartree-Fock + Bardeen-Cooper-Schrieffer (BCS) tables
Strength 4 Hartree-Fock-Bogoliubov (HFB) tables
Strength 5 Goriely’s hybrid model
Strength 6 Goriely Temperature-dependent HFB model
Strength 7 Temperature-dependent Relativistic Mean Field (RMF) model
Strength 8 Gogny D1M HFB + Quasiparticle Random Phase Approximation (QRPA)
Strength 9 Simplified Modified Lorentzian (SMLO) model
Show more

In this section, we select the 99Mo, 64Cu, and 67Cu radioisotopes as examples to compare the model calculations with the experimental data. 99mTc decayed by 99Mo is the most frequently used radioisotope in nuclear medicine. 68Zn is of interest owing to its applications in the production of the medical radioisotope 67Cu; 67Cu and 64Cu can form a “matched pair" [46, 47]. Therapeutic 67Cu, along with positron-emitting 64Cu, can measure the uptake kinetics in an organ of a patient by PET imaging, allowing for a precise dosimetric calculation. We used the default values of NLD and the γSF models in the following calculations.

3.1
99Mo/99mTc

With half-lives of 6 h and 140 keV γ-rays, 99mTc is nearly ideal for SPECT imaging. The most common method to obtain 99mTc is by elution from 99Mo generators. The 100Moγ, n reaction has a sufficient potential to produce 99Mo. The excitation functions are presented in Fig. 1, which covers the entire GDR energy range. The cross-sections of 100Moγ, n99Mo were experimentally measured by Utsunomiya et al. [48], Crasta et al. [49], and Ejiri et al. [14]. The gamma sources used in the experiments were LCS and bremsstrahlung. All the experimental results were distributed in the low-energy region of the GDR and did not exceed the peak. Fig. 1(a) demonstrates that the shape and value of the excitation function are highly sensitive to the choice of the γSF model, particularly at the peak position of GDR. Excluding strength 2 and 3, the remaining γSF models yield relatively similar results that are close to the experimental results. Different γSF models produce similar predictions for energies above the GDR peak. Overall, the cross-sections estimated using Strength 8 achieve the best agreement with the experimental data. As shown in Fig. 1(b), the calculation results of all the NLD models below the peak are consistent but significantly higher than the experimental data, whereas those above the peak are significantly different. The data predicted by the models at energies above 14 MeV are highly valuable. New experiments can be conducted based on SLEGS to verify the model predictions.

Fig. 1
(Color online) 100Mo(γ, n)99Mo reaction cross sections.
pic
3.2
64Cu

As listed in Table 1, the decay characteristics of 64Cu render it useful for nuclear medicine [50]. It combines PET diagnostic capabilities with those of radiotherapy, with average electron emissions of 190 keV. Currently, 64Cu is mainly produced by small cyclotrons through 64Ni(p,n) reactions [51-54]. Alternative production using 65Cuγ, n does not require rare or expensive 64Ni targets and simplifies the chemical separation step. The experimental data for the cross-sections of the 65Cuγ, n64Cu reaction are presented in Fig. 2 along with the TALYS. The measurements by Katz et al. [55] and Antonov et al. [56] were performed using bremsstrahlung gamma sources. However, they differed significantly in terms of their shape and value. Coote et al. [57] obtained a single cross-section at 17.6 MeV using monochromatic γ-rays from the 7Li(p, γ)8Be reaction. The cross-section is consistent with the calculation but lower than other experimental data by 80 to 120 mb. The cross-sections of 65Cuγ, n64Cu can significantly vary with the choice of the γSF models but not the NLD models. However, the discrepancies between the experiments and models cannot be compensated for by varying the γSF models. Future experiments regarding the 65Cu photoneutron reaction using a monochromatic LCS gamma source will be useful for resolving these discrepancies and provide assurance for its medical applications.

Fig. 2
(Color online) 65Cu(γ, n)64Cu reaction cross sections.
pic
3.3
67Cu

67Cu is a promising β--particle emitter with an average energy of 141 keV for the targeted radiotherapies. The size of these particles in the tissue was of the same order as the cell diameter. This reduces the unwanted dose burden on patients [58]. With a half-life of 61.83 h and low energy γ emissions (91.266 keV, 7%; 93.311 keV, 16.1%; 184.577 keV, 48.7%), 67Cu can provide a long therapeutic effect. 67Cu and its stable daughter 67Zn are nontoxic to the body, and both Cu and Zn are prevalent trace elements in the body. Along with the PET imaging radioisotope 64Cu, it forms a “matched pair”. However, the widespread clinical use of 67Cu-based radiopharmaceuticals has been limited by their availability, quantity, and quality. Experiments [59-63] have shown that the 68Zn(γ, p) reaction has the potential to produce sufficient quantities of 67Cu with a sufficient purity for medical use. Moreover, various Cu and Zn separation methods have been employed in radiochemical processing [34, 64-66].

Figure 3 presents the excitation function of the 68Zn(γ, p)67Cu reaction. Experimental data regarding this reaction are scarce, with only one group of two points in the energy region of SLEGS [67]. The data are significantly inconsistent with any calculation and hardly constrain NLD and γSF models. In contrast to the γ, n reaction of several hundred mb, the (γ, p) reaction is typically several mb in the medium-heavy nuclei. Additionally, protons can easily stop at the target and are difficult to detect. If the product nuclei of the (γ, p) reaction are unstable and have a moderate half-life, the cross-sections can be determined by offline measurements of the characteristic γ-rays during decay. For the 68Zn(γ, p) reaction, an offline measurement technique can be employed to determine its cross-sections.

Fig. 3
(Color online) 68Zn(γ, p)67Cu reaction cross sections.
pic
4

Summary

In this study, we summarize the medical radioisotopes that can be produced by γ, n, (γ, p), and (γ, γ′) reactions, along with the experimental data for these production pathways. We investigated the photonuclear reaction cross-sections for the production of 99Mo, 64Cu, and 67Cu. The experimental data for 99Mo and 67Cu did not cover the entire GDR energy region, whereas the data for 64Cu exhibited a significant discrepancy. To better constrain the NLD and γSF models in TALYS, additional experimental measurements at SLEGS in the energy region of GDR are necessary in the future.

To compare the productivity of the photonuclear method with that of traditional methods, we consider the production of 99Mo/99mTc as an example. The highest flux currently available for SLEGS is 107 s-1. We adopted the calculated results by the TALYS-1.96 code using the Strength 8 model for the yield calculations, which provided a maximum of 145 mb of the 100Mo(γ, n) reaction cross-section at the incident photon energy of 14.4 MeV. After irradiating a 100MoO3 target with a radius of 2 mm and a thickness of 1 cm for 5 times the half-life of 99Mo, the saturation specific activity of 99Mo and 99mTc can reach 4.76 and 4.74 μ Ci/g, respectively. The use of stacked targets can further enhance production. When the energy regions corresponding to the maximum cross-sections of the targeted nuclear reactions are close, multiple radioisotopes can be simultaneously produced. For example, the 62Cu, 64Cu, and 89Zr radioisotopes can be produced simultaneously when the incident photon energy is approximately 17 MeV.

Currently, 99Mo is primarily produced by the (n,f) reaction in a high-flux reactor using an enriched 235U target. The specific activity of 99Mo can reach 185 TBq/g (5000 Ci/g) [6]. However, most reactors will gradually shut down by 2030 [68]. Medical cyclotrons can be used to directly generate 99mTc via the 100Mo(p,2n) reaction. According to the experimental data, an enriched 100Mo target irradiated with a 16.5 MeV proton beam at 130 μA for 6 h yielded 116 GBq/g (3.13 Ci/g) of 99mTc [69]. Recently, a subcritical 99Mo production system was developed, which is driven by an accelerator-based deuterium–deuterium (D–D) neutron source. The D–D fusion reaction generates neutrons that irradiate a low-enriched uranium solution and induce fission in 235U. The system can generate 47.8 mCi/g 99Mo for a stable 24 h operation with a neutron intensity of 1×1014 n/s [70]. Bremsstrahlung gamma sources based on linear electron accelerators are often used to produce 99Mo radioisotopes. Using a 100MoO3 target irradiate with a 35 MeV electron beam at 100 μA for 20 h; the specific activity can achieve 4.4 GBq/g (119 mCi/g) [71]. NorthStar Medical Radioisotopes developed a 99Mo production system using an electron linear accelerator. Using this system based on γ, n reactions, approximately 30% more 99Mo is produced per gram of the target material compared with the traditional neutron capture route [72, 73].

In comparison, the specific activity produced by the photonuclear method based on SLEGS was low. With a further increase in the flux of γ beams, from 107 s-1 to 1015 s-1 [20], the specific activity of 99Mo can reach up to 4×102 Ci/g. For the 2-day protocol, the activity of the 99mTc-labelled tracers required for one myocardial perfusion imaging was 24 mCi [74]. The total yield of stacked targets in one year can provide 400,000 myocardial perfusion images. As the flux of SLEGS increases in the future, this method is promising for producing medical radioisotopes and will become feasible in China.

References
1. Ch. Schiepers(ed.) Diagnostic Nuclear Medicine (Springer & Berlin 2006).
2. G.J.R. Cook(ed.), Clinical Nuclear Medicine (Hodder Arnold, London, 2006).
3. R. Nutt,

The history of positron emission tomography

. Mol. Imaging Biol. 4, 11-26 (2002). https://doi.org/10.1016/S1095-0397(00)00051-0
Baidu ScholarGoogle Scholar
4. G.T. Gullberg, G.L. Zeng, F.L. Datz et al.,

Review of convergent beam tomography in single photon emission computed tomography

. Phys. Med. Biol. 37, 507-534 (1992). https://doi.org/10.1088/0031-9155/37/3/002
Baidu ScholarGoogle Scholar
5. J.L. Alberini, V. Edeline, A.L. Giraudet et al.,

Single photon emission tomography/computed tomography (SPET/CT) and positron emission tomography/computed tomography (PET/CT) to image cancer

. J. Surg. Oncol. 103, 602-606 (2011). https://doi.org/10.1002/jso.21695
Baidu ScholarGoogle Scholar
6.

World Nuclear Association, Radioisotopes in Medicine

, https://world-nuclear.org/information-library/non-power-nuclear-applications/radioisotopes-research/radioisotopes-in-medicine.aspx.Accessed 7 Feb 2023
Baidu ScholarGoogle Scholar
7.

IAEA: IAEA Annual Report for 2021

, https://www.iaea.org/opic/annual-report-2021
Baidu ScholarGoogle Scholar
8. D. Habs, U. Köster,

Production of medical radioisotopes with high specific activity in photonuclear reactions with γ-beams of high intensity and large brilliance

. Appl. Phys. B 103, 501-519 (2011). https://doi.org/10.1007/s00340-010-4278-1
Baidu ScholarGoogle Scholar
9. B. Szpunar, C. Rangacharyulu, S. Date et al.,

Estimates of production of medical isotopes by photo-neutron reaction at the Canadian Light Source

. Nucl. Instrum. Methods A 729, 41-50 (2013). https://doi.org/10.1016/j.nima.2013.06.106
Baidu ScholarGoogle Scholar
10. W. Luo, D.L. Balabanski, D. Filipescu,

A data-based photonuclear simulation algorithm for determining specific activity of medical radioisotopes

. Nucl. Sci. Tech. 27, 5 (2016). https://doi.org/10.1007/s41365-016-0111-9
Baidu ScholarGoogle Scholar
11. W. Luo, M. Bobeica, I. Gheorghe et al.,

Estimates for production of radioisotopes of medical interest at extreme light infrastructure–Nuclear physics facility

. Appl. Phys. B 122, 8 (2016). https://doi.org/10.1007/s00340-015-6292-9
Baidu ScholarGoogle Scholar
12. W. Luo,

Production of medical radioisotope 64Cu by photoneutron reaction using ELI-NP γ-ray beam

. Nucl. Sci. Tech. 27, 96 (2016). https://doi.org/10.1007/s41365-016-0094-6
Baidu ScholarGoogle Scholar
13. W.T. Pan, T. Song, H.Y. Lan et al.,

Photo-excitation production of medically interesting isomers using intense γ-ray source

. Appl. Radiat. Isot. 168, 109534 (2021). https://doi.org/10.1016/j.apradiso.2020.109534
Baidu ScholarGoogle Scholar
14. H. Ejiri, T. Shima, S. Miyamoto et al.,

Resonant photonuclear reactions for isotope transmutation

. J. Phys. Soc. Japan 80, 094202 (2011). https://doi.org/10.1143/JPSJ.80.094202
Baidu ScholarGoogle Scholar
15. J. Lee, H. Rehman, Y. Kim,

A feasibility study on the transmutation of 100Mo to 99mTc with laser-compton scattering photons

. Nucl. Technol. 201, 41-51 (2018). https://doi.org/10.1080/00295450.2017.1392397
Baidu ScholarGoogle Scholar
16. H.Y. Lan, D. Wu, J.X. Liu et al.,

Photonuclear production of nuclear isomers using bremsstrahlung induced by laser-wakefield electrons

. Nucl. Sci. Tech. 34, 74 (2023). https://doi.org/10.1007/s41365-023-01219-x
Baidu ScholarGoogle Scholar
17. IAEA(ed.) Non-HEU Production Technologies for Molybdenum-99 and Technetium-99m (IAEA, Vienna, 2013).
18. H.W. Wang, G.T. Fan, L.X. Liu et al.,

Commissioning of laser electron gamma beamline SLEGS at SSRF

. Nucl. Sci. Tech. 33, 87 (2022). https://doi.org/10.1007/s41365-022-01076-0
Baidu ScholarGoogle Scholar
19. Z.R. Hao, G.T. Fan, H.W. Wang et al.,

Collimator system of SLEGS beamline at Shanghai Light Source

. Nucl. Instr. Meth. A 1013, 165638 (2021). https://doi.org/10.1016/j.nima.2021.165638
Baidu ScholarGoogle Scholar
20. H.W. Wang, G.T. Fan, L.X. Liu et al.,

Development and prospect of Shanghai laser compton scattering gamma source

. Nucl. Phys. Rev. 37, 53-63 (2020). https://doi.org/10.11804/NuclPhysRev.37.2019043
Baidu ScholarGoogle Scholar
21. Z.R. Hao, G.T. Fan, L.X. Liu et al.,

Design and simulation of a 4π fat efciency 3He neutron detector array

. Nucl. Techn. 43, 110501 (2020). https://doi.org/10.11889/j.0253-3219.2020.hjs.43.110501
Baidu ScholarGoogle Scholar
22. K.J. Chen, L.X. Liu, Z.R. Hao et al.,

Simulation and test of the SLEGS TOF spectrometer at SSRF

. Nucl. Sci. Tech. 34, 47 (2023). https://doi.org/10.1007/s41365-023-01194-3
Baidu ScholarGoogle Scholar
23. P. Kuang, L.L Song, K.J. Chen et al.,

Nuclear resonance fluorescence spectrometer design and detector performance analysis of Shanghai laser electron gamma source(SLEGS)

. Nucl. Phys. Rev. 40, 58-65 (2023). https://doi.org/10.11804/NuclPhysRev.40.2022040
Baidu ScholarGoogle Scholar
24. Y.X. Yang, Y. Zhang, W.J. Zhao et al.,

Construction of gamma activation experimental platform for Shanghai laser electron gamma source

. Nucl. Phys. Rev. 31, 1-6 (2023). https://doi.org/10.11804/NuclPhysRev.31.01.01
Baidu ScholarGoogle Scholar
25. K. Goeke, J. Speth,

Theory of giant resonances

. Ann. Rev. Nucl. Part. Sci. 32, 65-115 (1982). https://doi.org/10.1146/ANNUREV.NS.32.120182.000433
Baidu ScholarGoogle Scholar
26. H. Feshbach,

Unified Theory of Nuclear Reactions

. Rev. Mod. Phys. 36, 1076 (1964). https://doi.org/10.1103/RevModPhys.36.1076
Baidu ScholarGoogle Scholar
27. M. Herman, G. Reffo, H.A. Weidenmüller,

Multistep-compound contribution to precompound reaction cross section

. Nucl. Phys. A 536, 124-140 (1992). https://doi.org/10.1016/0375-9474(92)90249-J
Baidu ScholarGoogle Scholar
28. L.Z. Dzhilavyan, A.I. Karev and V.G. Raevsky,

Possibilities for the production of radioisotopes for nuclear-medicine problems by means of photonuclear reactions

. Phys. Atom. Nuclei 74, 1690-1696 (2011). https://doi.org/10.1134/S1063778811120040
Baidu ScholarGoogle Scholar
29. M. Bobeica, D. Niculae, D. Balabanski et al.,

Radioisotopes production for medical applications at ELI-NP

. Rom. Rep. Phys. 68, S847-S883 (2016)
Baidu ScholarGoogle Scholar
30. A. Zilges, D.L. Balabanski, J. Isaak et al.,

Photonuclear reactions—From basic research to applications

. Prog. Part. Nucl. Phys. 122, 103903 (2022). https://doi.org/10.1016/j.ppnp.2021.103903
Baidu ScholarGoogle Scholar
31. D. Budker, J.C. Berengut, V.V. Flambaum et al.,

Expanding nuclear physics horizons with the gamma factory

. Ann. Phys. 534, 2100284 (2022). https://doi.org/10.1002/andp.202100284
Baidu ScholarGoogle Scholar
32.

Experimental Nuclear Reaction Data (EXFOR)

, https://www-nds.iaea.org/exfor/.Accessed 13 Feb 2023
Baidu ScholarGoogle Scholar
33.

NuDat 3 database

, https://www.nndc.bnl.gov/nudat3/.Accessed 13 Feb 2023
Baidu ScholarGoogle Scholar
34. R. Schwarzbach, K. Zimmermann, P. Bläuenstein et al.,

Development of a simple and selective separation of 67Cu from irradiated zinc for use in antibody labelling: A comparison of methods

. Appl. Radiat. Isot. 46, 329-336 (1995). https://doi.org/10.1016/0969-8043(95)00010-B
Baidu ScholarGoogle Scholar
35. M. Yagi, K. Kondo,

Preparation of carrier-free 47Sc by the 48Ti(γ, p)

. Int. J. Appl. Radiat. Isot. 28, 463-468 (1977). https://doi.org/10.1016/0020-708X(77)90178-8
Baidu ScholarGoogle Scholar
36. S.A. Kandil, B. Scholten, Z.A. Saleh et al.,

A comparative study on the separation of radiozirconium via ion-exchange and solvent extraction techniques, with particular reference to the production of 88Zr and 89Zr in proton induced reactions on yttrium

. J. Radioanal. Nucl. Chem. 274, 45-52 (2007). https://doi.org/10.1007/s10967-006-6892-2
Baidu ScholarGoogle Scholar
37. B.L. Fèvre, C. Pin,

A straightforward separation scheme for concomitant Lu–Hf and Sm–Nd isotope ratio and isotope dilution analysis

. Anal. Chim. Acta 543, 209-221 (2005). https://doi.org/10.1016/j.aca.2005.04.044
Baidu ScholarGoogle Scholar
38. V.N. Starovoitova, L. Tchelidze, D.P. Wells,

Production of medical radioisotopes with linear accelerators

. Appl. Radiat. Isot. 85, 39-44 (2014). https://doi.org/10.1016/j.apradiso.2013.11.122
Baidu ScholarGoogle Scholar
39. M. Inagaki, S. Sekimoto, W. Tanaka et al.,

Production of 47Sc, 67Cu, 68Ga, 105Rh, 177Lu, and 188Re using electron linear accelerator

. J. Radioanal. Nucl. Chem. 322, 1703-1709 (2019). https://doi.org/10.1007/s10967-019-06904-z
Baidu ScholarGoogle Scholar
40. A.G. Kazakov, T.Y. Ekatova, J.S. Babenya,

Photonuclear production of medical radiometals: a review of experimental studies

. J. Radioanal. Nucl. Chem. 328, 493-505 (2021). https://doi.org/10.1007/s10967-021-07683-2
Baidu ScholarGoogle Scholar
41. A. Koning, D. Rochman,

Modern nuclear data evaluation with the TALYS code system

. Nucl. Data Sheets 113, 2841-2934 (2012). https://doi.org/10.1016/j.nds.2012.11.002
Baidu ScholarGoogle Scholar
42. S. Stoulos, E. Vagena,

Indirect measurement of bremsstrahlung photons and photoneutrons cross sections of 204Pb and Sb isotopes compared with TALYS simulations

. Nucl. Phys. A 980, 1-14 (2018). https://doi.org/10.1016/j.nuclphysa.2018.09.081
Baidu ScholarGoogle Scholar
43. P.V. Cuong, T.D. Thiep, L.T. Anh et al.,

Theoretical calculation by Talys code in combination with Geant4 simulation for consideration of (γ, n) reactions of Eu isotopes in the giant dipole resonance region

. Nucl. Instrum. Methods B 479, 68-73 (2020). https://doi.org/10.1016/j.nimb.2020.06.011
Baidu ScholarGoogle Scholar
44. H. Cheng, B.H. Sun, L.H. Zhu et al.,

Measurements of 160Dy (p, γ) at energies relevant for the astrophysical γ process

. Astrophys. J. 915, 78 (2021). https://doi.org/10.3847/1538-4357/ac00b1
Baidu ScholarGoogle Scholar
45. W. Hauser, H. Feshbach,

The inelastic scattering of neutrons

. Phys. Rev. 87, 366-373 (1952). https://doi.org/10.1103/PhysRev.87.366
Baidu ScholarGoogle Scholar
46. A. Bockisch,

Matched pairs for radionuclide-based imaging and therapy

. Eur. J. Nucl. Med. Mol. Imaging 38, 1-3 (2011). https://doi.org/10.1007/s00259-011-1780-6
Baidu ScholarGoogle Scholar
47. S.M. Qaim, B. Scholten, B. Neumaier,

New developments in the production of theranostic pairs of radionuclides

. J. Radioanal. Nucl. Chem. 318, 1493-1509 (2018). https://doi.org/10.1007/s10967-018-6238-x
Baidu ScholarGoogle Scholar
48. H. Utsunomiya, S. Goriely, T. Kondo et al.,

Photoneutron cross sections for Mo isotopes: A step toward a unified understanding of (γ, n) and (n, γ) reactions

. Phys. Rev. C 88, 015805 (2013). https://doi.org/10.1103/PhysRevC.88.015805
Baidu ScholarGoogle Scholar
49. R. Crasta, H. Naik, S.V. Suryanarayana et al.,

Photo-neutron cross-section of 100Mo

. J. Radioanal. Nucl. Chem. 290, 367-373 (2011). https://doi.org/10.1007/s10967-011-1247-z
Baidu ScholarGoogle Scholar
50. C.J. Anderson, R. Ferdani,

Copper-64 radiopharmaceuticals for PET imaging of cancer: Advances in preclinical and clinical research

. Cancer Biother Radiopharm 24, 379-393 (2009). https://doi.org/10.1089/cbr.2009.0674
Baidu ScholarGoogle Scholar
51. J.Y. Kim, H. Park, J.C. Lee et al.,

A simple Cu-64 production and its application of Cu-64 ATSM

. Appl. Radiat. Isot. 67, 1190-1194 (2009). https://doi.org/10.1016/j.apradiso.2009.02.060
Baidu ScholarGoogle Scholar
52. Q.H. Xie, H. Zhu, F. Wang et al.,

Establishing reliable Cu-64 production process: From target plating to molecular specific tumor micro-PET imaging

. Molecules 22, 641 (2017). https://doi.org/10.3390/molecules22040641
Baidu ScholarGoogle Scholar
53. M. Jauregui-Osoro, S.D. Robertis, P. Halsted et al.,

Production of copper-64 using a hospital cyclotron: targetry, purification and quality analysis

. Nucl. Med. Commun. 42, 1024-1038 (2021). https://doi.org/10.1097/MNM.0000000000001422
Baidu ScholarGoogle Scholar
54. L. Isolan, M. Malinconico, W. Tieu et al.,

A digital twin for 64Cu production with cyclotron and solid target system

. Sci. Rep. 12, 19379 (2022). https://doi.org/10.1038/s41598-022-23048-5
Baidu ScholarGoogle Scholar
55. L. Katz, A.G.W. Cameron,

The solution of x-ray activation curves for photonuclear cross sections

. Can. J. Phys. 29, 518-544 (1951). https://doi.org/10.1139/p51-056
Baidu ScholarGoogle Scholar
56. A.D. Antonov, N.P. Balabanov, Yu.P. Gangrsky et al.

studied photonuclear reactions with the emission of α particles in the region of the Giant Dipole Resonance

. Yad. Fiz. 51, 305 (1990).
Baidu ScholarGoogle Scholar
57. G.E. Coote, W.E. Turchinetz, I.F. Wright,

Cross sections for the (γ, n) reaction in Cu63, Cu65, Zn64, Sb121 and Pr141, measured with monochromatic gamma rays

. Nucl. Phys. 23, 468-480 (1961). https://doi.org/10.1016/0029-5582(61)90273-5
Baidu ScholarGoogle Scholar
58. S.M. Qaim,

Therapeutic radionuclides and nuclear data

. Radiochim. Acta 89, 297-302 (2001). https://doi.org/10.1524/ract.2001.89.4-5.297
Baidu ScholarGoogle Scholar
59. V. Starovoitova, D. Foote, J. Harris et al.,

Cu‐67 photonuclear production

. AIP Conf. Proc. 1336, 502-504 (2011). https://doi.org/10.1063/1.3586150
Baidu ScholarGoogle Scholar
60. D.A. Ehst, N.A. Smith, D.L. Bowers et al.,

Copper-67 production on electron linacs—Photonuclear technology development

. AIP Conf. Proc. 1509, 157-161 (2012). https://doi.org/10.1063/1.4773959
Baidu ScholarGoogle Scholar
61. N.A. Smith, D.L. Bowers, D.A. Ehst,

The production, separation, and use of 67Cu for radioimmunotherapy: A review

. Appl. Radiat. Isot. 70, 2377-2383 (2012). https://doi.org/10.1016/j.apradiso.2012.07.009
Baidu ScholarGoogle Scholar
62. R.A. Aliev, S.S. Belyshev, A.A. Kuznetsov et al.,

Photonuclear production and radiochemical separation of medically relevant radionuclides: 67Cu

. J. Radioanal. Nucl. Chem. 321, 125-132 (2019). https://doi.org/10.1007/s10967-019-06576-9
Baidu ScholarGoogle Scholar
63. G.H. Hovhannisyan, T.M. Bakhshiyan, R.K. Dallakyan,

Photonuclear production of the medical isotope 67Cu

. Nucl. Instrum. Methods B 498, 48-51 (2021). https://doi.org/10.1016/j.nimb.2021.04.016
Baidu ScholarGoogle Scholar
64. N. Marceau, T.P.A. Kruck, D.B. McConnell et al.,

The production of copper 67 from natural zinc using a linear accelerator

. Int. J. Appl. Radiat. Isot. 21, 667-669 (1970). https://doi.org/10.1016/0020-708X(70)90121-3
Baidu ScholarGoogle Scholar
65. A.K. Dasgupta, L.F. Mausner, S.C. Srivastava,

A new separation procedure for 67Cu from proton irradiated Zn

. Appl. Radiat. Isot. 42, 371-376 (1991). https://doi.org/10.1016/0883-2889(91)90140-V
Baidu ScholarGoogle Scholar
66. M.A. Hassanein, H. El-Said, M.A. El-Amir,

Separation of carrier-free ^64,67Cu radionuclides from irradiated zinc targets using 6-tungstocerate(IV) gel matrix

. J. Radioanal. Nucl. Chem. 269, 75-80 (2006). https://doi.org/10.1007/s10967-006-0232-4
Baidu ScholarGoogle Scholar
67. A.E. Sioufi, P. Erdos, P. Stoll,

(γ, np) - Process in Mo-92 and Zn-66

. Helv. Phys. Acta 30, 264 (1957).
Baidu ScholarGoogle Scholar
68. Y.W. Wang, D.Y. Chen, R.S. Augusto et al.,

Production review of accelerator-based medical isotopes

. Molecules 27, 5294 (2022). https://doi.org/10.3390/molecules27165294
Baidu ScholarGoogle Scholar
69. P. Schaffer, F. Bénard, A. Bernstein et al.,

Direct production of 99mTc via 100Mo(p,2n) on small medical cyclotrons

. Phys. Procedia 66, 383-395 (2015). https://doi.org/10.1016/j.phpro.2015.05.048
Baidu ScholarGoogle Scholar
70. H.H. Xiong, Q.S. Zeng, Y.C. Han et al.,

Neutronics analysis of a subcritical blanket system driven by a gas dynamic trap-based fusion neutron source for 99Mo production

. Nucl. Sci. Tech. 34, 49 (2023). https://doi.org/10.1007/s41365-023-01206-2
Baidu ScholarGoogle Scholar
71. S. Sekimoto, K. Tatenuma, Y. Suzuki et al.,

Separation and purification of 99mTc from 99Mo produced by electron linear accelerator

. J. Radioanal. Nucl. Chem. 311, 1361-1366 (2017). https://doi.org/10.1007/s10967-016-4959-2
Baidu ScholarGoogle Scholar
72.

NorthStar Medical Radioisotopes: Accelerator Mo-99 Production

, https://www.northstarnm.com/development/accelerator-mo-99/.Accessed 15 Aug 2023
Baidu ScholarGoogle Scholar
73. J.T. Harvey, G.H. Isensee, G.P. Messina et al.,

Domestic production of Mo99

(2011), https://mo99.ne.anl.gov/2011/pdfs/Mo99%202011%20Web%20Presentations/S7-P3_Harvey.pdf.Accessed 15 Aug 2023
Baidu ScholarGoogle Scholar
74. B. Hesse, K. Tägil, A. Cuocolo et al.,

EANM/ESC procedural guidelines for myocardial perfusion imaging in nuclear cardiology

. Eur. J. Nucl. Med. Mol. Imaging 32, 855-897 (2005). https://doi.org/10.1007/s00259-005-1779-y
Baidu ScholarGoogle Scholar
75. B.C. Cook, A.S. Penfold, V.L. Telegdi,

Photodisintegration of C13

. Phys. Rev. 106, 300-314 (1957). https://doi.org/10.1103/PhysRev.106.300
Baidu ScholarGoogle Scholar
76. L.D. Cohen, W.E. Stephens,

Gamma-ray activation of carbon

. Phys. Rev. Lett. 2, 263-264 (1959). https://doi.org/10.1103/PhysRevLett.2.263
Baidu ScholarGoogle Scholar
77. J.P. Roalsvig, I.C. Gupta, R.N.H. Haslam,

Photoneutron reactions in C12 and O16

. Can. J. Phys. 39, 643-656 (1961). https://doi.org/10.1139/p61-075
Baidu ScholarGoogle Scholar
78. W.E.D. Bianco, W.E. Stephens,

Photonuclear activation by 20.5-MeV gamma rays

. Phys. Rev. 126, 709-717 (1962). https://doi.org/10.1103/PhysRev.126.709
Baidu ScholarGoogle Scholar
79. E.B. Bazhanov, A.P. Komar, A.V. Kulikov et al.,

Photodisintegration of C^12

. Yad. Fiz. 3, 711 (1966).
Baidu ScholarGoogle Scholar
80. B.C. Cook, J.E.E. Baglin, J.N. Bradford et al.,

C12(γ, n)C11 cross section to 65 MeV

. Phys. Rev. 143, 724-729 (1966). https://doi.org/10.1103/PhysRev.143.724
Baidu ScholarGoogle Scholar
81. W.A. Lochstet, W.E. Stephens,

12C(γ, n)11C giant resonance with gamma rays

. Phys. Rev. 141, 1002-1006 (1966). https://doi.org/10.1103/PhysRev.141.1002
Baidu ScholarGoogle Scholar
82. J.D. King, R.N.H. Haslam, R.W. Parsons,

The gamma-neutron cross section for N14

. Can. J. Phys. 38, 231-239 (1960). https://doi.org/10.1139/p60-023
Baidu ScholarGoogle Scholar
83. M.J. Facci, M.N. Thompson,

The absolute 14N(γ, n) reaction cross section

. Nucl. Phys. A 465, 77-82 (1987). https://doi.org/10.1016/0375-9474(87)90299-5
Baidu ScholarGoogle Scholar
84. B.C. Cook, J.E.E. Baglin, J.N. Bradford et al.,

O16(γ, n)O15 cross section from threshold to 65 MeV

. Phys. Rev. 143, 712-723 (1966). https://doi.org/10.1103/PhysRev.143.712
Baidu ScholarGoogle Scholar
85. B.S. Ishkhanov, I.M. Kapitonov, E.V. Lazutin et al.,

Fine structure of giant dipole resonance of the O16 nucleus

. Yad. Fiz. 12, 892 (1970).
Baidu ScholarGoogle Scholar
86. S.N. Belyaev, A.B. Kozin, A.A. Nechkin et al.,

Photoabsorption cross sections of Pb, Bi, and Ta isotopes in the energy region Eγ≤12 MeV

. Yad. Fiz. 42, 1050 (1985).
Baidu ScholarGoogle Scholar
87. S.N. Belyaev, V.A. Semenov,

Intermediate structure in (γ, n) cross sections at nuclei with N = 82

. Bull. Russ. Acad. Sci. Phys. 55, 66 (1991).
Baidu ScholarGoogle Scholar
88. P.R. Byerly, W.E. Stephens et al.,

Photodisintegration of copper

. Phys. Rev. 83, 54-62 (1951). https://doi.org/10.1103/PhysRev.83.54
Baidu ScholarGoogle Scholar
89. A.I. Berman, K.L. Brown,

Absolute cross section versus energy of the Cu63(γ, n) and Cu63(γ, 2n) reactions

. Phys. Rev. 96, 83-89 (1954). https://doi.org/10.1103/PhysRev.96.83
Baidu ScholarGoogle Scholar
90. M.B. Scott, A.O. Hanson, D.W. Kerst,

Electro- and photodisintegration cross sections of Cu63

. Phys. Rev. 100, 209-214 (1954). https://doi.org/10.1103/PhysRev.100.209
Baidu ScholarGoogle Scholar
91. T. Nakamura, K. Takamatsu, K. Fukunaga et al.,

Absolute cross sections of the (γ, n) reaction for Cu63, Zn64, and Ag109

. J. Phys. Soc. Japan 14, 693-698 (1959). https://doi.org/10.1143/JPSJ.14.693
Baidu ScholarGoogle Scholar
92. S. Yasumi, M. Yata, K. Takamatsu et al.,

Absolute cross section of the reaction Cu63(γ, n)Cu62 for Lithium gamma rays

. J. Phys. Soc. Japan 15, 1913-1919 (1960). https://doi.org/10.1143/JPSJ.15.1913
Baidu ScholarGoogle Scholar
93. R.E. Sund, M.P. Baker, L.A. Kull et al.,

Measurements of the 63Cu(γ, n) and (γ, 2n) cross sections

. Phys. Rev. 176, 1366-1376 (1968). https://doi.org/10.1103/PhysRev.176.1366
Baidu ScholarGoogle Scholar
94. D.G. Owen, E.G. Muirhead, B.M. Spicer,

Structure in the giant resonance of 64Zn and 63Cu

. Nucl. Phys. A 122, 177-183 (1968). https://doi.org/10.1016/0375-9474(68)90711-2
Baidu ScholarGoogle Scholar
95. L.Z. Dzhilavyan, N.P. Kucher,

Measurement of the cross section of the reaction 63Cu(γ, n) in a beam of quasimonochromatic annihilation photons in the energy region 12-25 MeV

. Yad. Fiz. 30, 294 (1979).
Baidu ScholarGoogle Scholar
96. S.C. Jeong, J.C. Kim, B.N. Sung,

Cross section measurement for the 63Cu(γ, n)62Cu reaction

. J. Korean Phys. Soc. 17, 359 (1984).
Baidu ScholarGoogle Scholar
97. M.N. Martins, E. Hayward, G. Lamaze et al.,

Experimental test of the bremsstrahlung cross section

. Phys. Rev. C 30, 1855-1860 (1984). https://doi.org/10.1103/PhysRevC.30.1855
Baidu ScholarGoogle Scholar
98. C. Plaisir, F. Hannachi, F. Gobet et al.,

Measurement of the 85Rb(γ, n)84mRb cross-section in the energy range 10-19 MeV with bremsstrahlung photons

. Eur. Phys. J. A 48, 68 (2012). https://doi.org/10.1140/epja/i2012-12068-7
Baidu ScholarGoogle Scholar
99. K. Masumoto, T. Kato, N. Suzuki,

Activation yield curves of photonuclear reactions for multielement photon activation analysis

. Nucl. Instrum. Methods 157, 567-577 (1978). https://doi.org/10.1016/0029-554X(78)90019-8
Baidu ScholarGoogle Scholar
100. A.K.M.L. Rahman, K. Kato, H. Arima et al.,

Study on effective average (γ, n) cross section for 89Y, 90Zr, 93Nb, and 133Cs and (γ, 3n) cross section for 99Tc

. J. Nucl. Sci. Technol. 47, 618-625 (2010). https://doi.org/10.3327/jnst.47.618
Baidu ScholarGoogle Scholar
101. A. Banu, E.G. Meekins, J.A. Silano et al.,

Photoneutron reaction cross section measurements on 94Mo and 90Zr relevant to the p-process nucleosynthesis

. Phys. Rev. C 99, 025802 (2019). https://doi.org/10.1103/PhysRevC.99.025802
Baidu ScholarGoogle Scholar
102. T. Shizuma, H. Utsunomiya, P. Mohr et al.,

Photodisintegration cross section measurements on 186W, 187Re, and 188Os: Implications for the Re-Os cosmochronology

. Phys. Rev. C 72, 025808 (2005). https://doi.org/10.1103/PhysRevC.72.025808
Baidu ScholarGoogle Scholar
103. V.A. Zheltonozhsky, M.V. Zheltonozhskaya, A.M. Savrasov et al.,

Studying the activation of 177Lu in (γ, pxn) reactions

. Bull. Russ. Acad. Sci. Phys. 84, 923-928 (2020). https://doi.org/10.3103/S1062873820080328
Baidu ScholarGoogle Scholar
104. O.V. Bogdankevich, L.E. Lazareva, A.M. Moiseev,

Inelastic scattering on 103Rh nuclei

. J. Exp. Theor. Phys. 12, 853 (1961).
Baidu ScholarGoogle Scholar
105. Z.M. Bigan, V.A. Zheltonozhsky, V.I. Kirishchuk et al.,

Excitation of 113In, 195Pt and 199Hg isomers in reactions of inelastic gamma scattering

. Bull. Russ. Acad. Sci. Phys. 70, 292 (2006).
Baidu ScholarGoogle Scholar
106. V.G. Nedorezov, E.S. Konobeevski, S.V. Zuyev et al.,

Excitation of 111mCd, 113mIn, and 115mIn isomeric states by photons of energy up to 8 MeV

. Phys. Atom. Nucl. 80, 827-830 (2017). https://doi.org/10.1134/S1063778817050167
Baidu ScholarGoogle Scholar
107. O.V. Bogdankevich, L.E. Lazareva, F.A. Nikolaev,

Inelastic scattering of photons by indium-115 nuclei

. J. Exp. Theor. Phys. 4, 320 (1957).
Baidu ScholarGoogle Scholar
108. V.M. Mazur, I.V. Sokolyuk, Z.M. Bigan et al.,

Cross section for excitation of nuclear isomers in (γ, γ’)-m reactions at 4-15 MeV

. Yad. Fiz. 56, 20 (1993).
Baidu ScholarGoogle Scholar
109. N.A. Demekhina, A.S. Danagulyan, G.S. Karapetyan,

Formation of isomeric states in (γ, γ’) reactions at energies around the giant dipole resonance

. Phys. Atom. Nucl. 64, 1796-1798 (2001). https://doi.org/10.1134/1.1414927
Baidu ScholarGoogle Scholar
110. V.S. Bokhinyuk, A.I. Guthy, A.M. Parlag et al.,

Study of the effective excitation cross section of the 115In isomeric state in the (γ, γ’) reaction

. Ukr. J. Phys. 51, 657 (2006).
Baidu ScholarGoogle Scholar
111. W. Tornow, M. Bhike, S.W. Finch et al.,

Measurement of the 115In(γ, γ’)115mIn inelastic scattering cross section in the 1.8 to 3.7 MeV energy range with monoenergetic photon beams

. Phys. Rev. C 98, 064305 (2018). https://doi.org/10.1103/PhysRevC.98.064305
Baidu ScholarGoogle Scholar
Footnote

The authors declare that they have no competing interests.