logo

A substitution measurement for cross section of 65Cu(γ, n)64Cu reaction using natCu and 63Cu targets by quasi-monoenergetic γ beams at SLEGS

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

A substitution measurement for cross section of 65Cu(γ, n)64Cu reaction using natCu and 63Cu targets by quasi-monoenergetic γ beams at SLEGS

Pu Jiao
Zi-Rui Hao
Zhi-Cai Li
Qian-Kun Sun
Long-Xiang Liu
Hang-Hua Xu
Yue Zhang
Meng-Die Zhou
Wen Luo
Yu-Xuan Yang
Sheng Jin
Kai-Jie Chen
Shan Ye
Zhen-Wei Wang
Yu-Ting Wang
Hui-Ling Wei
Yao Fu
Kun Yu
Hong-Wei Wang
Gong-Tao Fan
Chun-Wang Ma
Nuclear Science and TechniquesVol.36, No.12Article number 240Published in print Dec 2025Available online 17 Oct 2025
15004

To overcome the difficulty and high cost of some specific isotopic targets, a substitution method was proposed to measure the cross section of the (γ, n) reactions. Considering that the natural copper element (natCu) only has 63Cu and 65Cu isotopes, the 65Cu(γ, n)64Cu reaction was taken as an example to test the substitution method. Using quasi-monoenergetic γ beams provided by the Shanghai Laser Electron Gamma Source (SLEGS) of the Shanghai Synchrotron Radiation Facility (SSRF), natCu(γ, n) was measured from Eγ=11.09 MeV to 17.87 MeV. Furthermore, based on the 63Cu(γ, n) reaction measured using the same experimental setup at SLEGS, 65Cu(γ, n)64Cu was extracted using the substitution method. The abundance variation of natural copper, showing a significant influence on the cross section, was also investigated. The results were compared to the existing experimental data measured by bremsstrahlung and positron annihilation in-flight sources, and the TALYS 2.0 predictions. The γ strength function (γSF) of 65Cu was obtained from the 65Cu(γ, n) data, and the reaction cross section of 64Cu(n, γ) was further calculated.

Photoneutron cross sectionFlat efficiency detectorLaser Compton scatteringγ raysSLEGSSubstitution measurementCopper isotopes
1

Introduction

The 65Cu(γ, n)64Cu reaction has important medical and scientific applications [1, 2]. 64Cu is a short-life β+ emitter (T1/2=12.7 h; β+ with a mean energy of 278 keV, a branching ratio of 61.5%, and β- with a mean energy of 191 keV and a branching ratio of 38.5%), which is widely used in nuclear medical imaging techniques such as positron emission tomography (PET) [3, 4] and single-photon emission computed tomography (SPECT) [5] and plays an important role in clinical diagnosis. For example, 64Cu-labeled peptides such as 64Cu-DOTATATE [1, 6] are used in the diagnosis of neuroendocrine tumors [7], 64Cu-labeled oxygen depletion probes, such as 64Cu-ATSM [8], are used to detect oxygen depletion regions in tumors, and 64Cu-labeled prostate specific membrane antigen (PSMA) [9] ligands are used in PET imaging of prostate cancer to precisely localize tumor cells. In addition, 64Cu can be used in radionuclide therapy to destroy tumor cells through the decay properties of β+ and β- [10]. The 65Cu(γ, n)64Cu reaction is also of great importance in scientific research. For example, in nuclear physics research, the reaction allows the study of the structure and properties of atomic nuclei, as well as nuclear reaction mechanisms. The photoneutron reaction to produce 64Cu has advantages over traditional methods, such as avoiding the use of rare and expensive 64Ni targets and complex chemical separation. In addition, the 64Cu(n, γ)65Cu reaction plays a key role in the quality control of the medical isotope 64Cu. It helps assess the potential loss of 64Cu during neutron irradiation and provides valuable data for understanding the nucleosynthesis of elements of medium mass in stars through neutron capture processes.

In the last century, laboratories worldwide have conducted experimental studies on the photoneutron reaction of 65Cu using bremsstrahlung (BR) sources [11, 12] and positron annihilation in flight (PAIF) sources [13] using the activation method. Varlamov et al. [14] evaluated the existing 65Cu(γ,n) experimental data, which showed considerable differences. The evaluated and experimental cross sections show that BR and PAIF are close within the low incident γ energy range but differ significantly in the high γ energy range, reflecting systematic errors in the experiment caused by the misclassification of neutron channels [8]. The quasimonochromatic γ-ray source generated by laser Compton scattering provides an opportunity to measure the (γ, n) reaction, which helps distinguish the differences in the existing data.

In this study, the cross sections of the natCu(γ, n) reaction were measured within the giant dipole resonance (GDR) energy region using the SLEGS beamline [15] at the SSRF [16-18]. The 65Cu(γ, n) cross sections were determined via the substitute method via the previously measured 63Cu(γ, n) reaction. Furthermore, neutron capture cross sections for 64Cu were also extracted. The remainder of this paper is organized as follows. Section 2 describes the experimental procedure for the natCu(γ, n) cross sections. Section 3 presents the methods for processing the experimental data of the photoneutron cross section and the results of the quasi-monochromatic and monochromatic cross sections of 65Cu(γ, n) obtained by the subtraction method. Section 4 discusses the discrepancies between the measured data and existing experimental data, as well as the extraction of the radiative neutron capture cross section of 64Cu. Finally, a brief conclusion is given in Sect. 5.

2

Experiment

This experiment was performed at the SLEGS beamline station [19] in the SSRF. The beamline uses inverse Compton scattering technology: 3.5 GeV electrons in the SSRF storage ring collide with photons from a 10.64 μm-wavelength, 100 W CO2 laser, generating quasimonochromatic gamma rays with tunable energies from 0.66 to 21.7 MeV. The energy of the γ beam was adjusted in slant-scattering mode with a minimum step of 10 keV. For measurements of the (γ, n) reactions at SLEGS, see Refs. [20-23] for details. A schematic illustration of SLEGS and the corresponding experimental setup are presented in Fig. 1.

Fig. 1
(Color online) Schematic layout of the SLEGS beamline
pic

The cross sections for the natCu(γ, n) reaction were measured at 44 energy points ranging from 10.9 MeV (θ = 90°) to 17.8 MeV (θ = 130°). For each angle, the measurement time and neutron statistics were as follows: 2 h with neutron statistics exceeding 1.0 × 104 for θ ≤ 98°, 1 h with neutron statistics exceeding 4.3 × 104 for 99° ≤ θ ≤ 150°, and 0.5 h with neutron statistics exceeding 4.8×104 for θ ≤ 106°. After passing through the collimation system, the laser Compton scattering (LCS) γ beam irradiated the experimental target positioned at the center of the 3He flat efficiency detector (FED) array [24]. The in-beam gamma flux was monitored using a large-volume BGO detector downstream of the FED. The incident γ spectrum was reconstructed using the direct unfolding method combined with a Geant4-simulated detector-response matrix (Fig. 2, see Refs. [25-27] for details).

Fig. 2
(Color online) The energy spectrum of incident γ ray beams in the experiments
pic
2.1
Targets

The natCu target (3.15 g) was placed in polyethylene target holders and irradiated by LCS γ beams. The alignments of the target and FED with the LCS γ-ray beam were adjusted using a MiniPIX X-ray pixel detector for collimation. The detailed specifications are provided in Table 1.

Table 1
The information of natCu targets used in experiments
Target Weight (g) Diameter (mm) Total thickness (mm) Measured density (g/cm3)
natCu 3.15 10.00 4.48 8.69
Show more
Total chemical purity: natCu > 99.99%

The target holder has a 10-mm in diameter window. Considering that the size of the LCS γ-ray beams was approximately 4 mm in diameter at the target position, a 10 mm diameter window was sufficient for the target to be measured, avoiding the influence of neutrons from polythene.

2.2
Measurements

The SLEGS facility features a new FED with 26 proportional counters arranged in three concentric radii within a polyethylene moderator shielded by a 2 mm Cd sheet [23]. The counters, with an effective length of 500 mm and filled with 3He gas at a pressure of 2 atm, were read out through the Mesytec MDPP-16 digitizers and MVME DAQ. Figure 3 shows the efficiency curves of each ring and the total efficiency curve simulated by GEANT4 using a real detector configuration. For the neutron evaporation spectrum, the total detector efficiency increases from 35.64% at 50 keV to 42.32% at 1.65 MeV, and then decreases slowly to 39.05% at 4 MeV [28]. The efficiency calibrated using the 252Cf source is 42.10 ± 1.25%, corresponding to an average neutron energy of 2.13 MeV. In our experiment, we used the ring-ratio technique to obtain the average energy of neutrons produced by the (γ, n) reaction and then estimated the detector efficiency using its calibration curve [29, 30].

Fig. 3
(Color online) Total detector efficiency and efficiencies of individual rings. Detector efficiency curves were simulated using neutron evaporation spectra and monochromatic neutrons. The red dots are given by the neutron spectrum described by the Maxwell-Boltzmann distribution, at the average neutron energy (T = 1.42 MeV) of 252Cf [28]
pic
3

Data Analysis and Substitution Measurement Method

3.1
Data Analysis method

In the monochromatic approximation, the photoneutron cross section can be expressed by the integral equation [31]: SnEmaxnγ(E)σ(E)dE=NnNγNtξϵn, (1) where nγ(E) is the energy distribution of the LCS γ-ray beam normalized in the integration region. σ(E) represents the photoneutron cross section, Nn the number of detected neutrons, Nt the number of target nuclei per unit area; and Nγ the number of incident γ-ray with energies above the neutron threshold. The self-attenuation coefficient ξ is given by [32] ξ=μt1eμt, (2) where μ is the linear attenuation coefficient of the sample and t is the thickness of the sample. The photoneutron cross section in the monochromatic approximation is calculated by σ(γ,n)Emax=NnNγNtξϵn. (3) In the experiment, the laser pulse cycle was 1000 μs (50 μs on + 950 μs off). Owing to the energy dispersion of LCS γ-ray beams, the monochromatic approximation is insufficient for determining photoneutron cross sections. When neutrons were counted with the FED array, the flat-efficiency regions for each detector ring were determined, considering the neutron energy and detector parameters. The median method was used to establish the optimal efficiency points and improve neutron count statistics.

To solve this unfolding problem, the integral in Eq. (1) is approximated as the summation of each γ beam profile, resulting in a system of linear equations σf=Dσ [33, 34]. The folding iteration method was used to solve this underdetermined system. Starting with a constant trial function σ0, the folded vector σf0=Dσ0 was calculated [35-38]. The next trial input function σ1 was obtained by adding the difference between the experimental spectrum σexp and the folded spectrum σf0 to Dσ0 after spline interpolation to match the vector dimensions. The iteration proceeds with [39-41] σi+1=σi+(σexpσfi). (4) The iteration continues until convergence, when σfi+1 approximates σexp within statistical errors. Convergence was assessed by calculating the reduced χ2 between σfi+1 and σexp, with typical convergence achieved in approximately three iterations, resulting in a reduced χ2 value of approximately 1.

3.2
Substitution Measurement method

In this experiment, the natural Cu target (natCu) has an isotopic abundance of 69.15% for 63Cu and 30.85% for 65Cu. The one-neutron (Sn) and two-neutrons (S2n) separation energies for 63Cu are 10.86 and 19.74 MeV, respectively. And for 65Cu, Sn and S2n become 9.91 MeV and 17.83 MeV, respectively [42-44]. Within the energy range where the one-neutron separation energy thresholds of 63Cu and 65Cu overlap, the FED detector measures neutrons from both the 63Cu(γ, n) and 65Cu(γ, n) reactions as NnatCu=N63Cu+N65Cu, (5) where NnatCu represents the result obtained from direct measurement, whereas N63Cu needs to be calculated by combining the monoenergy cross section σ(γ,n)63Cu with the current energy spectrum nγ(Eγ); the detailed calculation process is shown in Eq. (6), where d is the thickness of target. N63Cu=NtgϵnSnEmaxnγdEγ(1eμnatCuρnatCud)μnatCuρnatCud ×SnEmax nγ(Eγ)σ(γ,n)63Cu(Eγ)dEγ. (6) By substituting the neutron count of N63Cu into Eq. (7), the quasi-monoenergetic photoneutron cross section data of 65Cu can be obtained as σ(γ,n)65Cu=SnEmax nγ(E)σ(E)dE=NnatCuN63CuNγNtξϵn. (7) The monochromatic cross section of the 65Cu(γ, n)64Cu reaction was derived using a deconvolution iteration method. Figure 4 compares the quasi-monoenergetic and monochromatic cross sections of 65Cu.

Fig. 4
(Color online) Cross sections of 65Cu(γ, n) measured at SLEGS. The dots are the folded cross section and the line with shaded area is the unfolded (monochromatic) cross section
pic

According to Eq. (7), statistical uncertainty is mainly caused by Nn. Methodological uncertainty arises from the extraction algorithm Nn and deconvolution method incorporating the simulated BGO response matrix. Systematic uncertainty, which is the main source of the total error, includes the γ-flux uncertainty from the copper attenuator, the uncertainty of the target thickness, and the uncertainty of the FED efficiency. Table 2 summarizes the systematic and methodological uncertainties.

Table 2
Summary of (γ,n) cross section uncertainties for measurements at SLEGS
Error Source Type Uncertainty
FED efficiency Systematic 3.02%
External copper Systematic 0.50%
Target thickness Systematic <0.10%
Nn extraction algorithm Data processing 2.00%
Unfolding method Data processing 1.00%
Show more
4

Results and discussion

4.1
65Cu photoneutron reaction cross section

The 65Cu(γ,n)64Cu photoneutron cross section data measured at SLEGS were compared with the existing experimental and evaluated data in Fig. 5. Although the overall trends were consistent, significant discrepancies were observed in the absolute values. The experiment by Fultz et al. [13] at Lawrence Livermore National Laboratory (LLNL) is closest to the results in this work. They used BF3 neutron detectors and measured in the energy range of 9.34 to 27.78 MeV. Data from Katz et al. [11] using BR source at the BETAT accelerator in Canada with a 22 MeV endpoint energy, are notably higher than the LLNL data. Antonov’s [12] measurements using the BR source at JINR are significantly higher than those of other datasets in terms of both neutron threshold and cross-sectional values. The cross section for 65Cu(γ, n)64Cu reaction measured by the monoenergetic LCS-γ source at SLEGS is listed in Table 3. Data errors include statistical, methodological, and systematic uncertainties. The total uncertainties at different are also listed in Table 3.

Fig. 5
(Color online) Cross section of 65Cu(γ, n)64Cu reaction. The solid circles denote the measured results with the natural abundance (30.85%) of 65Cu. The results measured using BR γ rays by Katz [11] and Antonov [12] are plotted as solid green diamonds and purple squares, respectively. The results measured using PAIF γ rays by Fultz [13] are plotted as solid blue triangles. The TENDL-2021 evaluation is indicated by a sky-blue solid line. The calculated results by the increased (+5.00%) and decreased (-5.00%) abundance of 65Cu are indicated by the purple line segment and gray dotted line. The inset figure highlights the differences in cross sections at the peak positions for these isotopic abundances
pic
Table 3
The monochromatic cross sections of 65Cu(γ, n)64Cu measured at SLEGS by substitution method
Eγ (MeV) Cross sections (mb) Statistical uncertainty (mb) Systematical uncertainty (mb) Methodological uncertainty (mb) Total uncertainty (mb)
11.09 11.38 0.26 0.08 0.16 0.42
11.28 12.80 0.22 0.12 0.21 0.49
11.47 14.13 0.21 0.13 0.23 0.58
11.66 15.39 0.19 0.17 0.30 0.67
11.85 16.67 0.19 0.23 0.40 0.76
12.03 18.07 0.19 0.26 0.45 0.85
12.22 19.69 0.20 0.30 0.51 0.94
12.41 21.54 0.19 0.33 0.59 1.03
12.60 23.61 0.28 0.38 0.66 1.12
12.78 25.84 0.31 0.40 0.71 1.20
12.97 28.29 0.31 0.46 0.79 1.29
13.16 31.04 0.33 0.50 0.82 1.41
13.34 34.21 0.32 0.54 0.89 1.54
13.53 37.85 0.33 0.57 0.95 1.70
13.71 41.91 0.29 0.65 1.06 1.88
13.89 46.31 0.42 0.71 1.18 2.07
14.07 50.93 0.43 0.78 1.30 2.27
14.25 55.79 0.54 0.85 1.37 2.48
14.43 60.93 0.56 0.98 1.61 2.70
14.61 66.59 0.45 1.01 1.62 2.95
14.79 72.81 0.52 1.19 1.94 3.25
14.96 79.64 0.47 1.16 1.93 3.58
15.14 86.86 0.51 1.34 2.21 3.95
15.31 94.16 0.50 1.45 2.43 4.34
15.48 101.13 0.53 1.59 2.64 4.71
15.66 107.55 0.51 1.65 2.75 5.07
15.82 113.23 0.63 1.93 3.20 5.40
15.99 118.06 0.66 2.06 3.40 5.70
16.16 122.15 0.68 2.20 3.62 5.97
16.32 125.43 0.70 2.33 3.84 6.20
16.65 129.18 0.60 2.57 4.26 6.51
16.96 128.71 0.60 2.78 4.77 6.58
17.27 123.81 0.68 2.84 4.85 6.39
17.58 115.61 0.82 2.97 4.97 6.01
17.87 105.68 0.81 2.98 5.03 5.52
Show more

Isotopic abundance variations influence cross sections and cause discrepancies in the results. The photoneutron cross sections changed regularly with alterations in 65Cu abundance. Based on data from the 2025 SLEGS experiment, comparing the cross section of 65Cu at 35.85%, natural abundance (30.85%), and 25.85%, the photoneutron cross sections decreased as the abundance of 65Cu increased (e.g., from 25.85% to 35.85%) and increased as the abundance decreased (see the inset in Fig. 5). This occurs across all energy ranges and is most noticeable near the threshold and peak positions in the cross section distribution. It is suggested that the isotopic abundances in the target material should be determined prior to analysis of the substitution data. The sensitive change in isotopic abundance also suggests that it is capable of adopting an enhanced isotopic target to determine the cross section (γ, n) in addition to the pure isotopic target.

As discussed in Ref. [45], the ratios of the integral cross sections provide a clear indication of the systematic differences among the various data compilations. The integral cross sections in Sn and Smax regions are as follows: σint=SnSmaxσ(E)dE. (8) Based on these experimental data, the integral ratios of the photoneutron reaction cross section were calculated for energy ranges from Sn to 15 MeV, 15 MeV to S2n, and Sn to S2n, and the results are presented in Table 4. In the energy range of Sn to 15 MeV, the measured results show a difference of only 0.8% from the Fultz data and a discrepancy of less than 0.4% from theoretical calculations of TENDL-2021, while discrepancies with other datasets exceed 40%. In the 15 MeV to S2n range, the minimum difference between this study and the Katz [11] data is 0.4%, and the differences with other datasets exceed 20%. In general, the neutron threshold and the peak position of the cross section demonstrate good consistency with the results measured by Fultz [13] and the evaluation in TENDL-2021 [46]. The neutron threshold exhibits favorable agreement with Katz data [11]. However, there are notable differences in both the neutron threshold and peak position compared with the data measured by Antonov [12].

Table 4
Integral cross section ratio
Ratio relation σint ratio
  Sn~15 MeV 15 MeV ~S2n Sn~S2n
σTENDLint/σSLEGSint 1.04 0.68 0.80
σFultzint/σSLEGSint 0.92 0.61 0.69
σKatzint/σSLEGSint 1.49 1.04 1.21
σAntonovint/σSLEGSint 0.23 0.77 0.68
Show more
4.2
64Cu radiative neutron capture cross section

The gamma strength function (γSF) [47] is used to describe the average probabilities of gamma decay and absorption in nuclear reactions, and is an important parameter for characterizing the nuclear reaction process. When research involves reactions with gamma rays, such as reactions (n, γ) and (γ, n), the precision of γSF is particularly crucial. According to the principle of detailed balance [48] and generalized Brink assumption [49-51], it is believed that the upward fXl(Eγ) is approximately equal to the downward fXl(Eγ). Therefore, it can be considered that fXl(Eγ)fXl(Eγ)fXl(Eγ). The (upward) σγn photoneutron cross section is connected to the (downward) γSF [52] by fX1(Eγ)=13π22c2σγn(Eγ)Eγ, (9) in which the constant is 1/3π22c2=8.674×108 mb-1MeV-2. Using this relationship, γSF can be obtained from the cross section measured in the photoneutron reaction. Then, the γSF model in TALYS was compared with the experimentally constrained γSF. The model prediction closest to the experimentally obtained γSF is selected. Furthermore, the γSF model is constrained using the normalization parameter Gnorm in the TALYS 2.0 toolkit, and the optimization of the normalization parameter Gnorm is achieved by minimizing the χ2 value, thus making the theoretical calculation results of the constrained γSF model more consistent with the experimentally obtained γSF values. The γSF of 65Cu(γ, n) constrained by measured (γ, n) data is shown in Fig. 6, with χ2 determined by χ2=1Ni=1N(σth,iσexp,iσerr,i)2, (10) where N denotes the total number of experimental data points. σth, i, σexp, i and σerr, i denote the theoretical value, experimentally measured value, and experimental error of γSF at the i-th data point, respectively. The neutron capture cross section of 64Cu, after adjustment of the optimal Gnorm value, is shown in Fig. 7. Specifically, Gnorm is set to 1.2 in the Kopecky-Uhl generalized Lorentzian model [53]. However, it is 1.4 in the Goriely hybrid model. In particular, when constrained by Gnorm, the χ2 value of the hybrid mode [54] was the smallest among all the investigated models, showing the best agreement with this set of experimental data. Similarly, Utsunomiya et al. [39] and Li et al. [44] previously conducted related research and measured the (n, γ) radiative reaction cross sections for the 136,137Ba and 62Cu isotopes. In this study, owing to the lack of experimental data on the low-lying excited states and neutron resonance spacings of 65Cu, the constraints on the nuclear level density (NLD) model are limited, resulting in substantial theoretical uncertainties. The study of the 64Cu(n,γ) cross section is of great value for improving nuclear data and optimizing the preparation of medical isotopes. This is closely related to the measurement of the 65Cu(γ,n)64Cu cross section through the principle of detailed balance, and the former can verify the reliability of the latter, collectively highlighting the overall significance of the research.

Fig. 6
(Color online) The γSF values of 65Cu calculated using the Kopecky-Uhl generalized Lorentzian model (red solid line) and Goriely’s hybrid model (blue dash-dot line) with default parameters are compared with those obtained from the optimized Kopecky-Uhl generalized Lorentzian model (red dashed line) and Goriely’s hybrid model (blue dotted line) using the Gnorm method, along with the γSF values extracted from the SLEGS experimental data (red plot points)
pic
Fig. 7
(Color online) The cross sections for 64Cu(n, γ)65Cu are calculated using the optimized Opecky-Uhl generalized Lorentzian model (red shaded area)and Goriely’s hybrid model (purple shaded area). The shaded area represents the results considering six different nuclear-level density models within the TALYS 2.0 [51]
pic
5

Summary

The reaction cross sections of natCu(γ, n) were measured in the incident energy range of 11.09 to 17.87 MeV using the 3He FED detector array developed by SLEGS. Based on the measured photoneutron cross section data and the previously measured results for 63Cu(γ, n) at SLEGS, the reaction cross sections of 65Cu(γ, n)64Cu were obtained using the cross section substitution method. Compared with existing experimental data, the reliability of this method was demonstrated, providing a new approach for photoneutron cross section measurements. Given the extensive application of 64Cu in medical fields such as nuclear medicine imaging and tumor therapy, clarifying the existing discrepancies in the reaction cross-sections of the 65Cu(γ, n)64Cu reaction is likely to play an important role in these fields. The sensitivity of the isotopic abundance change in the natural copper target shows that the enhanced purity of the specific isotope could be used to measure its (γ, n) cross section via the substitution measured in this work. The experimentally constrained γSF of 65Cu was extracted from the 65Cu(γ, n)64Cu cross section distribution. In addition, the cross section curve of its inverse reaction, 64Cu(n, γ), was calculated, which provides a new approach to the extraction of cross sections (n, γ) from some unstable nuclides.

References
1. 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
2. W. Luo, D. L. Balabanski, D. Filipescu,

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

, Nucl. Sci. Tech. 27, 113 (2016). https://doi.org/10.1007/s41365-016-0111-9.
Baidu ScholarGoogle Scholar
3. I. M. Jackson, P. J. H. Scott, S. Thompson,

Clinical applications of radiolabeled peptides for PET

. Semin. Nucl. Med. 47, 493 (2017). https://doi.org/10.1053/j.semnuclmed.2017.05.007.
Baidu ScholarGoogle Scholar
4. M. Brandt, J. Cardinale, M. L. Aulsebrook et al.,

An overview of PET radiochemistry, part 2: Radiometals

. J. Nucl. Med. 59, 1500 (2018). https://doi.org/10.2967/jnumed.117.190801.
Baidu ScholarGoogle Scholar
5. O. O. Krasnovskaya, D. Abramchuck, A. Erofeev et al.,

Recent advances in 64Cu/67Cu-based radiopharmaceuticals

. Int. J. Mol. Sci. 24, 9154 (2023). https://doi.org/10.3390/ijms24119154.
Baidu ScholarGoogle Scholar
6. M. Bobeica, D. Niculae, D. Balabanski et al.,

Radioisotope production for medical applications at ELI-NP

. Rom. Rep. Phys. 68, 847 (2016).
Baidu ScholarGoogle Scholar
7. A. Boschi, P. Martini, E. Janevik-Ivanovska et al.,

The emerging role of copper-64 radiopharmaceuticals as cancer theranostics

. Drug Discov. Today 23, 1489 (2018). https://doi.org/10.1016/j.drudis.2018.04.002.
Baidu ScholarGoogle Scholar
8. F. Cantiello, V. Gangemi, G. L. Cascini et al.,

Diagnostic accuracy of 64copper prostate-specific membrane antigen positron emission tomography/computed tomography for primary lymph node staging of intermediate-to high-risk prostate cancer: our preliminary experience

. Urology. 106, 139 (2017). https://doi.org/10.1016/j.urology.2017.04.019.
Baidu ScholarGoogle Scholar
9. B. Grubmüller, R. P. Baum, E. Capasso et al.,

64Cu-PSMA-617 PET/CT imaging of prostate adenocarcinoma: first in-human studies

. Cancer Biother. Radiopharm. 31, 277 (2016). https://doi.org/10.1089/cbr.2015.1964.
Baidu ScholarGoogle Scholar
10. G. A. Follacchio, M. S. De Feo, G. De Vincentis et al.,

Radiopharmaceuticals labelled with copper radionuclides: clinical results in human beings

. Curr. Radiopharm. 11, 22 (2018). https://doi.org/10.2174/1874471011666171211161851.
Baidu ScholarGoogle Scholar
11. H. E. Johns, L. Katz, R. E. A. Douglas et al.,

Gamma-neutron cross sections

. Phys. Rev. 80, 1062 (1950). https://doi.org/10.1103/PhysRev.80.1062.
Baidu ScholarGoogle Scholar
12. A. D. Antonov, N. P. Balabanov, Yu. P. Gangrskii et al.,

Studies of photonuclear reactions with emission of α particles in the region of the giant dipole resonance

. Sov. J. Nucl. Phys. 51, 193 (1990).
Baidu ScholarGoogle Scholar
13. S. C. Fultz, R. L. Bramblett, J. T. Caldwell et al.,

Photoneutron cross sections for natural Cu, Cu63, and Cu65

. Phys. Rev. 133, B1149 (1964). https://doi.org/10.1103/PhysRev.133.B1149.
Baidu ScholarGoogle Scholar
14. A. V. Varlamov, V. V. Varlamov, D. S. Rudenko et al., Atlas of giant dipole resonances. Parameters and graphs of photonuclear reaction cross sections. INDC (NDS)-394, IAEA NDS, Vienna, Austria. 1-311(1999). https://inis.iaea.org/records/4exm3-2ag10.
15. L. X. Liu, H. W. Wang, G. T. Fan et al.,

The SLEGS beamline of SSRF

, Nucl. Sci. Tech. 35, 111 (2024). https://doi.org/10.1007/s41365-024-01469-3
Baidu ScholarGoogle Scholar
16. R. Z. Tai, Z. T. Zhao,

Overview of SSRF phase-II beamlines

, Nucl. Sci. Tech. 35, 137 (2024). https://doi.org/10.1007/s41365-024-01487-1
Baidu ScholarGoogle Scholar
17. R. Z. Tai, Z. T. Zhao,

Commissioning and First Results of the SSRF Phase-II Beamline Project

, J. Phys. Conf. Ser. 2380, 012004 (2022). https://doi.org/10.1088/1742-6596/2380/1/012004
Baidu ScholarGoogle Scholar
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. 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(3), 47 (2023). https://doi.org/10.1007/s41365-023-01194-3
Baidu ScholarGoogle Scholar
20. H. H. Xu, G. T. Fan, H. W. Wang et al.,

Interaction chamber for laser Compton slant-scattering in SLEGS beamline at Shanghai Light Source

, Nucl. Instrum. Methods phys. Res. A 1033, 166742 (2022). https://doi.org/10.1016/j.nima.2022.166742
Baidu ScholarGoogle Scholar
21. Z. R. Hao, H. H. Xu, G. T. Fan et al.,

Gamma spot monitor at SLEGS beamline

, Nucl. Instrum. Methods Phys. Res. A 1068, 169748 (2024). https://doi.org/10.1016/j.nima.2024.169748
Baidu ScholarGoogle Scholar
22. Z. C. Li, Y. X. Yang, Z. W. Cao et al.,

Effective extraction of photoneutron cross-section distribution using gamma activation and reaction yield ratio method

, Nucl. Sci. Tech. 34, 170 (2023). https://doi.org/10.1007/s41365-023-01330-z
Baidu ScholarGoogle Scholar
23. M. D. Zhou, G. T. Fan, C. W. Ma et al.,

Measurement of 59Co(γ, n)58Co using a new flat-efffciency neutron detector at SLEGS

, Phys. Rev. C 111, 054612 (2025). https://doi.org/10.1103/PhysRevC.111.054612
Baidu ScholarGoogle Scholar
24. Z. R. Hao, L. X. Liu, Y. Zhang et al.,

The day-one experiment at SLEGS: systematic measurement of the (γ, 1n) cross sections on 197Au and 159Tb for resolving existing data discrepancies

, Science Bulletin (2025). https://doi.org/10.1016/j.scib.2025.05.037
Baidu ScholarGoogle Scholar
25. L. X. Liu, H. Utsunomiya, G. T. Fan et al.,

Energy profile of laser Compton slant-scattering γ-ray beams determined by direct unfolding of total-energy responses of a BGO detector

, Nucl. Instrum. Methods Phys. Res. A 1063, 169314 (2024). https://doi.org/10.1016/j.nima.2024.169314
Baidu ScholarGoogle Scholar
26. W. Luo, H. Y. Lan, Y. Xu et al.,

Implementation of the n-body Monte-Carlo event generator into the Geant4 toolkit for photonuclear studies

, Nucl. Instrum. Methods Phys. Res. A 849, 49 (2017). https://doi.org/10.1016/j.nima.2017.01.010
Baidu ScholarGoogle Scholar
27. S. Agostinelli, J. Allison, K. A. Amako et al.,

Geant4—a simulation toolkit

, Nucl. Instrum. Methods Phys. Res. A 506, 250 (2003). https://doi.org/10.1016/S0168-9002(03)01368-8
Baidu ScholarGoogle Scholar
28. Z. R. Hao, G. T. Fan, L. X. Liu et al.,

Design and simulation of 4π flat-efficiency 3He neutron detector array

, Nucl. Tech. (in Chinese) 43, 9 (2020). https://doi.org/10.11889/j.0253-3219.2020.hjs.43.110501
Baidu ScholarGoogle Scholar
29. B. Berman, J. Caldwell, R. Harvey et al.,

Photoneutron cross sections for 90Zr, 91Zr, 92Zr, 94Zr, and 89Y

, Phys. Rev. 162, 1098 (1967). https://doi.org/10.1103/PhysRev.162.1098
Baidu ScholarGoogle Scholar
30. O. Itoh, H. Utsunomiya, H. Akimune et al.,

Photoneutron cross sections for Au revisited: measurements with laser Compton scattering γ-rays and data reduction by a least-squares method

, J. Nucl. Sci. Technol. 48, 834 (2011). https://doi.org/10.1080/18811248.2011.9711766
Baidu ScholarGoogle Scholar
31. J. W. Jury, B. L. Berman, D. D. Faul et al.,

Photoneutron cross sections for 13C

. Phys. Rev. C 19, 1684 (1979). https://doi.org/10.1103/PhysRevC.19.1684
Baidu ScholarGoogle Scholar
32. National Institute of Standards and Technology (NIST), X-ray mass attenuation coefficients. NIST X-ray Mass Attenuation Coefficients (2025)
33. Y. L. Dang, L. F. Long, F. G. Yong et al.,

New measurement of thick target yield for narrow resonance at Ex=9.17 MeV in the 13C(p,γ)14N reaction

. Chin. Phys. B 28, 060706 (2019). https://doi.org/10.1088/1674-1056/ab96a0
Baidu ScholarGoogle Scholar
34. F. L. Liu, W. S. Yang, J. H. Wei et al.,

Study on γ-ray source from the resonant reaction 19F(p,αγ)16O at Ep=340 keV

. Chin. Phys. B 29, 070702 (2020). https://doi.org/10.1088/1674-1056/ab96a0
Baidu ScholarGoogle Scholar
35. M. Guttormsen, T. S. Tveter, L. Bergholt et al.,

The unfolding of continuum γ-ray spectra

. Nucl. Instrum. Methods Phys. Res. A 374, 3 (1996). https://doi.org/10.1016/0168-9002(96)00197-0
Baidu ScholarGoogle Scholar
36. H. Utsunomiya, I. Gheorghe, D. M. Filipescu et al.,

Direct neutron-multiplicity sorting with a flat-efficiency detector

. Nucl. Instrum. Methods Phys. Res. A 871, 135 (2017). https://doi.org/10.1016/j.nima.2017.08.001
Baidu ScholarGoogle Scholar
37. I. Gheorghe, H. Utsunomiya, S. Katayama et al.,

Photoneutron cross-section measurements in the 209Bi(γ,xn) reaction with a new method of direct neutron-multiplicity sorting

. Phys. Rev. C 96, 044604 (2017). https://doi.org/10.1103/PhysRevC.96.044604
Baidu ScholarGoogle Scholar
38. I. Gheorghe, H. Utsunomiya, K. Stopani et al.,

Updated neutron-multiplicity sorting method for producing photoneutron average energies and resolving multiple firing events

. Nucl. Instrum. Methods Phys. Res. A 1019, 165867 (2021). https://doi.org/10.1016/j.nima.2021.165867
Baidu ScholarGoogle Scholar
39. T. Renstrøm, H. Utsunomiya, H. T. Nyhus et al.,

Verification of detailed balance for γ absorption and emission in Dy isotopes

, Phys. Rev. C 98, 054310 (2018). https://doi.org/10.1103/PhysRevC.98.054310
Baidu ScholarGoogle Scholar
40. G. C. Yang, L. M. Hua, F. Lu et al.,

Response functions of a 4 π summing gamma detector in β-Olso method

, Nucl. Sci. Tech. 33, 68 (2022). https://doi.org/10.1007/s41365-022-01058-2
Baidu ScholarGoogle Scholar
41. P. Jiao, Z. R. Hao, Q. K. Sun et al.,

Measurements of 27Al(γ, n) reaction using quasi-monoenergetic γ beams from 13.2 to 21.7 MeV at SLEGS

. Nucl. Sci. Tech. 36, 66 (2025). https://doi.org/10.1007/s41365-025-01662-y
Baidu ScholarGoogle Scholar
42. W. J. Huang, M. Wang, F. G. Kondev et al.,

The AME 2020 atomic mass evaluation (I). Evaluation of input data, and adjustment procedures

, Chin. Phys. C 45, 030002 (2021) https://doi.org/10.1088/1674-1137/abddb0
Baidu ScholarGoogle Scholar
43. M. Wang, W. J. Huang, F. G. Kondev et al.,

The AME 2020 atomic mass evaluation (II). Tables, graphs and references

, Chin. Phys. C 45 030003, (2021). https://doi.org/10.1088/1674-1137/abddaf
Baidu ScholarGoogle Scholar
44. Z. C. Li, Z. R. Hao, Q. K. Sun et al.,

New measurement of 63Cu(γ,n)62Cu cross-section using quasi-monoenergetic γ-ray beam

. Nucl. Sci. Tech. 36, 34 (2025). https://doi.org/10.1007/s41365-024-01631-x
Baidu ScholarGoogle Scholar
45. V. Varlamov, B. Ishkhanov, V. Orlin,

Reliability of (γ, 1n),(γ, 2n), and (γ, 3n) cross-section data on 159Tb

, Phys. Rev. C 95, 054607 (2017). https://doi.org/10.1103/PhysRevC.95.054607
Baidu ScholarGoogle Scholar
46.

TENDL-2021

, https://tendl.web.psi.ch/tendl_2021/gamma_html/gamma.html, (2021).
Baidu ScholarGoogle Scholar
47. J. Kopecky, M. Uhl,

Test of gamma-ray strength functions in nuclear reaction model calculations

. Phys. Rev. C 41.5, 1941 (1990). https://doi.org/10.1103/PhysRevC.41.1941
Baidu ScholarGoogle Scholar
48. J. M. Blatt, V. F. Weisskopf, F. J. Dyson et al., Theoretical nuclear physics. American Institute of Physics (1953). https://doi.org/10.1063/1.3061164
49. H. Utsunomiya, T. Renstrϕm, G. M. Tveten et al.,

Photoneutron cross sections for Ni isotopes: Toward understanding (n,γ) cross sections relevant to weak s-process nucleosynthesis

. Phys. Rev. C 98, 054619 (2018). https://doi.org/10.1103/PhysRevC.98.054619
Baidu ScholarGoogle Scholar
50. D. Filipescu, I. Gheorghe, H. Utsunomiya et al.,

Photoneutron cross sections for samarium isotopes: Toward a unified understanding of (γ,n) and (n,γ) reactions in the rare earth region

. Phys. Rev. C 90, 064616 (2014). https://doi.org/10.1103/PhysRevC.90.064616
Baidu ScholarGoogle Scholar
51. A. Koning,

TEFAL-2.0, Making ENDF-6 nuclear data libraries with TALYS

, https://nds.iaea.org/talys/tutorials/tefal.pdf (2023)
Baidu ScholarGoogle Scholar
52. R. Capote, M. Herman, P. Obložinskỳ et al.,

RIPL-reference input parameter library for calculation of nuclear reactions and nuclear data evaluations

. Nucl. Data Sheets. 110, 3107 (2009). https://doi.org/10.1016/j.nds.2009.10.004
Baidu ScholarGoogle Scholar
53. S. Goriely, V. Plujko.

Simple empirical E1 and M1 strength functions for practical applications

. Phys. Rev. C 99, 014303 (2019). https://doi.org/10.1103/PhysRevC.99.014303
Baidu ScholarGoogle Scholar
54. E. B. Balbutsev, I. V. Molodtsova, P. Schuck.

Spin scissors mode and the fine structure of M1 states in nuclei

. Nucl. Phys. A 872, 1 (2011). https://doi.org/10.1016/j.nuclphysa.2011.09.016
Baidu ScholarGoogle Scholar
Footnote

Chun-Wang Ma and Hong-Wei Wang are editorial board members for Nuclear Science and Techniques and was not involved in the editorial review, or the decision to publish this article. All authors declare that there are no competing interests.