logo

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

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

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

Zhi-Cai Li
Zi-Rui Hao
Qian-Kun Sun
Yu-Long Chen
Long-Xiang Liu
Hang-Hua Xu
Yue Zhang
Pu Jiao
Meng-Die Zhou
Yu-Xuan Yang
Sheng Jin
Kai-Jie Chen
Zhen-Wei Wang
Shan Ye
Xin-Xiang Li
Chun-Wang Ma
Hong-Wei Wang
Gong-Tao Fan
Wen Luo
Nuclear Science and TechniquesVol.36, No.2Article number 34Published in print Feb 2025Available online 14 Jan 2025
16006

We present new data on the 63Cu(γ, n) cross section studied using a quasi-monochromatic and energy-tunable γ beam produced at the Shanghai Laser Electron Gamma Source to resolve the long-standing discrepancy between existing measurements and evaluations of this cross section. Using an unfolding iteration method, 63Cu(γ, n) data were obtained with an uncertainty of less than 4%, and the inconsistencies between the available experimental data were discussed. The γ-ray strength function of 63Cu(γ, n) was successfully extracted as an experimental constraint. We further calculated the cross-section of the radiative neutron capture reaction 62Cu(n, γ) using the TALYS code. Our calculation method enables the extraction of (n, γ) cross sections for unstable nuclides.

63Cu(γ, n) reactionCross section dataQuasi-monochromatic γ beamRadiative neutron capture reaction
1

Introduction

62Cu, which can be produced by the 63Cu(γ, n) reaction, is a relatively short-lived β+ emitter (T1/2 = 9.67 min) suitable for positron emission tomography (PET) imaging. For example, [62Cu]Cu para-toluene sulfonic acid methyl ester (PTSM) provides high quality brain and heart images with PET, accurately delineating cerebral and myocardial perfusion in both animals and humans [1]. Accurate cross-sectional data for the 63Cu(γ, n) reaction are required to guide the production of the 62Cu isotope for medical purposes [2, 3]. Moreover,the 63Cu(γ, n) reaction cross-section data can be applied to monitor the bremsstrahlung radiation [4] and LCS γ-ray [5] fluxes. In addition, the β+ decay process of 62Cu is an alternative pathway for the synthesis of the “iron group element” 62Ni, and knowledge of the radiative neutron capture reaction 62Cu(n, γ) can help us understand the nucleosynthesis of intermediate-mass elements, as shown in Fig. 1.

Fig. 1
(Color online) Nuclear reaction path around the Cu isotope
pic

Over the past few decades, 63Cu photoneutron reactions have been experimentally studied worldwide using electron-accelerator-based bremsstrahlung radiation or positron annihilation in flight γ beam facilities [4, 6-11]. Cross-sectional data of 63Cu(γ, n) and 63Cu(γ, 2n) reactions were obtained and subsequently evaluated by Varlamov et al. [12, 13]. However, discrepancies in the shapes, peak heights, and positions of these experimental and evaluated cross-sectional data curves were observed. Additionally, the peak widths of these curves were different [14]. These discrepancies must be resolved. For example, the experimental data of Fultz [7] were systematically 15% lower than those of Varlamov [12]. A recent article [15] reported significantly different peak values of 79.79 mb and 59±6 mb for the evaluated 63Cu(γ, n) and experimental data, respectively [7]. Varlamov et al. [13] analyzed experimental 63Cu(γ, 2n) data [7] using an experimental-theoretical procedure [16, 17], indicating the need for reasonable verification and correction of these data. Luo et al. [18] proposed a method for extracting the (γ, n) cross-sectional distribution of ~40 isotopes, including 63Cu, using laser-induced γ activation and the isotope yield ratio.

Furthermore, owing to the extreme difficulty in obtaining the 62Cu target, experimental data on the 62Cu(n, γ) reaction are unavailable in any neutron energy range [14]. However, such data can be extracted indirectly from the cross section of its inverse 63Cu(γ, n) reaction, which requires accurate 63Cu(γ, n) data with sufficiently small measurement uncertainty. Consequently, obtaining new data on the 63Cu(γ, n) reaction and calculating the cross-section of the 62Cu(n, γ) reaction are essential.

The Shanghai Laser Electron Gamma Source (SLEGS) is an energy-tunable laser Compton scattering (LCS) γ-ray source that provides MeV γ beams for nuclear science and technology [19-21]. It was developed based on the inverse Compton scattering of 10.64 μm CO2 laser photons from 3.5 GeV relativistic electrons in the storage ring of the Shanghai Synchrotron Radiation Facility (SSRF) [22, 23]. SLEGS delivers γ beams with energies of 0.66–21.1 MeV in the slant-scattering mode at scattering angles of 20°–160° and at a maximum energy of 21.7 MeV in the back-scattering mode at 180°. The full-spectrum flux ranges from ~105 photons/s at 20° to ~107 photons/s at 180° [24, 25]. SLEGS provides a suitable experimental platform for conducting various types of photonuclear reaction experiments and is particularly suitable for experimental measurements of the (γ, n) and (γ, 2n) cross sections in the giant dipole resonance (GDR) energy region.

In this study, we experimentally investigated the 63Cu(γ, n) cross section based on quasi-monochromatic and energy-tunable SLEGS γ beams. Using an unfolding iteration method, 63Cu(γ, n) data were obtained within the energy range of 11.1–19.7 MeV. Then, the γ-ray strength function (γSF) of the 63Cu(γ, n) reaction was extracted, and the cross-section of its inverse reaction, 62Cu(n, γ), was successfully calculated. The remainder of this paper is organized as follows. In Sect. 2, the experimental procedure used to measure the 63Cu(γ, n) cross section is described. In Sect. 3, the results of the monochromatic and unfolded 63Cu(γ, n) cross sections are presented. In Sect. 4, the inconsistency between the available experimental data from different laboratories is discussed, and the experimentally constrained γSF for the 63Cu(γ, n) reaction and cross-sectional data of the inverse reaction, 62Cu(n, γ), are presented. Finally, a brief conclusion is presented in Sect. 5.

2

Experimental Procedure

The experimental measurement of the 63Cu(γ, n) cross section was conducted at SLEGS of the SSRF, which produces γ beams within the GDR energy range from the single neutron separation energy (Sn=10.86 MeV) to the double neutron separation energy (S2n=19.74 MeV). The experimental setup is schematically illustrated in Fig. 2. An energy-tunable SLEGS γ beam was generated through a slanting LCS process, which was achieved by the interaction of the SSRF electron beam and a CO2 laser with incident angles ranging from 20° to 160°. After collimation, the quasi-monochromatic γ-ray was guided to irradiate the 63Cu target, which was positioned precisely at the geometric center of a 3He flat efficiency detector (FED) array. During the experiments, the neutrons produced from the photoneutron processes were first moderated by polyethylene in the FED array and then detected by 3He proportional counters. The γ beam penetrating the target was attenuated by an additional copper attenuator (naturally abundant), and its spectrum was subsequently measured using a bismuth germanate (BGO) detector.

Fig. 2
(Color online) Schematic of the experimental setup for measuring the photoneutron cross section
pic
2.1
SLEGS γ beam Spectrum

In parallel with the 63Cu irradiation, the γ-ray spectrum after copper attenuation was measured online using a BGO detector. Figure 3 shows an exemplary γ-ray spectrum detected at the slant scattering angles (θL) of 91°, 113°, and 140°. To obtain the SLEGS γ-ray spectrum in front of the irradiation target, a direct unfolding method was employed in combination with a known response function of the BGO detector, which was obtained by a GEANT4 simulation [26]. The resulting (unfolded) γ-ray spectrum is shown in Fig. 3. The folded-back spectrum is consistent with the γ-ray spectrum measured by the detector, suggesting reliable reproduction of the γ-ray spectrum before the irradiation target. The γ-ray spectrum was integrated and corrected for the Cu attenuation factor to obtain the γ beam flux at each slant scattering angle.

Fig. 3
(Color online) Typical γ-ray spectra measured by the BGO detector (black line), folded-back γ-ray spectra (red line), and corresponding γ-ray spectra incident on the target (blue line) at θL = 91°, 113°, and 140°
pic
2.2
63Cu Target

The diameter, thickness, and purity of the 63Cu target were a 10 mm, 1.5 mm, and 99.8%, respectively. For the 63Cu isotope sample, the purity of the 63Cu target was determined by inductively coupled plasma-mass spectrometry. The uncertainty of the thickness was estimated to be 0.01 mm.

2.3
Neutron Detection

The number of (γ, n) reactions was determined by detecting the reaction neutrons using the calibrated FED, which comprised 26 sets of 3He proportional counters embedded in a polyethylene moderator. The proportional counters were arranged in three concentric rings positioned 65 mm, 110 mm, and 175 mm from the beam axis. All the sensitive volumes of the 3He proportional counters were cylindrical in shape with the same length of 500 mm and inflated with 3He gas at 2 atm. While the counters in Ring-1 (inner ring) were 1 inch in diameter, those in Ring-2 (middle ring) and Ring-3 (outer ring) were 2 inches in diameter. The bodies of 3He proportional counters were made of stainless steel for a lower α emission rate. The inner polyethylene moderator was 450 mm × 450 mm × 550 mm (along the beam direction) and was surrounded by additional polyethylene plates with cadmium to suppress the background neutrons [27]. These background neutrons were subtracted from the duty cycle of the laser pulse. In our experiments, the duty cycle is set to 50 μs per laser period of 1000 μs. The polyethylene moderation effect significantly broadened the time distribution of the neutrons detected by the 3He proportional counters. However, a flat interval of the time distribution that was only contributed by background neutrons was identifiable. Then, the number of neutrons (Nn) was directly extracted by subtracting the time-normalized background. Further details are available in [28].

The average energy of the reaction neutrons was obtained using the “ring-ratio technique” originally developed by Berman and Fultz [29] and used to determine the detection efficiency. Fig. 4(a) shows the simulated efficiency curve by GEANT4 with realistic detector configuration. For the neutron-evaporation spectra, the total detector efficiency increases from 35.64% at 50 keV to 42.32% at 1.65 MeV and then falls slowly to 39.05% at 4 MeV. The efficiency calibrated using a 252Cf source was 42.10 ± 1.25%, corresponding to an average neutron energy of 2.13 MeV. In our experiments, we used the ring-ratio technique to obtain the average energy of neutrons produced by (γ, n) reactions and then estimated the detector efficiency using its calibrated curve of the detector efficiency. The curve for the efficiency ratio of Ring-3 to Ring-1 is illustrated in Fig. 4(b) [28].

Fig. 4
(Color online) (a) Total detector efficiency and the efficiencies of individual rings. The detector efficiency curves were simulated by neutron-evaporation spectra. The red dots are given by the neutron spectrum described by the Maxwell-Boltzmann distribution, P(E)E1/2exp(E/T), at the average neutron energy (T = 1.42 MeV) of 252Cf. (b) Ring-ratio curve of the FED
pic
3

Data Analysis and Results

3.1
Monochromatic Cross Section

The experimental formula for the photoneutron cross section is given by [30, 31] SnEmaxnγ(Eγ)σ(Eγ)dEγ=NnNγNtξϵng, (1) where () is the spectral distribution of the normalized LCS γ beam; σ() is the photoneutron cross section; Nn is the number of neutrons detected; Nt is the number of target nuclei per unit area; is the number of γ-rays incident on the target; ϵn is the neutron detection efficiency; and ξ=(1-eμd)/μd is a correction factor for a thick-target measurement. Here, μ is the linear attenuation coefficient of γ photons in a target of thickness d. The factor g represents the fraction of γ flux above the neutron threshold Sn: g=SnEmaxnγ(Eγ)dEγ0Emaxnγ(Eγ)dEγ. (2) The incident γ energy distribution was used to determine the cross section σ(), which is a function of the γ energy . Specifically, the incoming γ beam spectra were used to determine (). The γ-energy distribution was normalized to unity: SnEmaxnγ(Eγ)dEγ=1. The measured σexpEmax for an incoming γ beam with maximum energy Emax is given by the convoluted cross-section: σexpEmax=SnEmaxnγ(Eγ)σ(Eγ)dEγ=NnNγNtξϵng. (3) Therefore, we refer to the quantity on the right side of Eq. (3) as the monochromatic cross section. However, owing to the energy spread of the LCS γ beam (Fig. 3), the monochromatic approximation cannot be used to describe the real photoneutron cross section.

3.2
Unfolded Cross Section

The deconvoluted -dependent photoneutron cross-section, σ(), must be extracted from the integral of Eq. (3). Each measurement characterized by Emax corresponds to the folding of σ() with the measured beam profile nγ(Eγ). Following Ref. [32], we unfold σ() according to Eq. (3): σf=Dσ, (4) where σf represents the folded cross section with beam profile D. The indices i and j of matrix element Dij correspond to Emax and , respectively. The set of equations is given by: [σ1σ2σN]f=[D11D12D1MD21D22D2MDN1DN2DNM][σ1σ2σM]. (5) Each row of D corresponds to the γ beam profile corresponding to Emax. The σ vector [σi]f (i=1, 2, 3..., N) on the left-hand side of Eq. (5) is the folded cross-section, referred to as the experimental monochromatic cross-section, whereas the vector [σj] (j=1,2,3..., M) on the right-hand side is the unfolded cross-section to be determined. In the present experiment, the number of monochromatic cross sections was N=44. The energy profile of the γ beam was simulated in M=200 energy bins. The number of unfolded cross-sections was equal to M. Figure 3 presents a visual representation of the response matrix D for the case of 63Cu. There are N=44 γ-beam spectra, and only three are shown as blue lines in Fig. 3 as examples. As the system of linear equations in Eq. (5) is under-determined, the σj vector cannot be obtained by matrix inversion. We determined σj using an iterative folding method, which can be summarized as follows:

(1) As our starting point, we choose a constant trial function σ0 for the zeroth iteration. This initial vector is multiplied by D to obtain the zeroth folded vector σf0=Dσ0,

(2) The next trial input function, σ1, is established by adding the difference between the experimentally measured spectrum σexp and folded spectrum σf0 to σ0. To add the folded and input vectors, we first perform a spline interpolation on the folded vector, and then interpolate to ensure that the two vectors have equal dimensions. The new input vector is σ1=σ0+(σexpσf0). (6) (3) The above steps are iterated i times, yielding σfi=Dσi, (7) and σi+1=σi+(σexpσfi), (8) until convergence is achieved. Thus, σfi+1σexp is within the statistical uncertainties. To check the convergence quantitatively, we calculated the reduced χ2 of σfi+1 and σexp after each iteration. The experiment was terminated when the reduced χ2 value approached unity.

Figure 5 shows the monochromatic σexpEmax and unfolded σ() for 63Cu(γ, n) reaction. Table 1 lists the σ() values at Emax and their uncertainties respectively. According to Eq. (3), the statistical uncertainty is primarily induced by Nn. Because the incident γ-ray count was sufficiently high, its statistical uncertainty was negligible. The methodological uncertainty was approximately 1.8%, which was induced by the extraction algorithm Nn (1.5%) and unfolding methodology incorporating the simulated BGO response matrix (~1%). The systematic uncertainty was estimated to be 3.15%. This was due to the neutron detector efficiency (3.02%), γ flux attenuation and incident γ spectrum unfolding (0.90%), and target areal density (0.10%). In our study, the total uncertainty included statistical, systematic, and methodological uncertainties. The unfolded σ() had a total uncertainty of approximately 4%, except for the γ-energy region with σ() less than 7.5 mb (i.e., Sn < < 11.5 MeV).

Fig. 5
(Color online) 63Cu(γ, n) reaction cross section as a function of the incident γ energy, . The dots indicate the monochromatic cross section, while the line with the shadow area indicates the unfolded cross section
pic
Table 1
Unfolded cross sections and corresponding uncertainties for 63Cu(γ, n)62Cu
(MeV) σ (mb) Statistical uncertainty (mb) Methodological uncertainty (mb) Systematic uncertainty (mb) Total uncertainty (mb)
11.09 2.18 0.40 0.03 0.06 0.41
11.28 4.70 0.23 0.05 0.08 0.25
11.47 7.28 0.13 0.06 0.10 0.17
11.66 9.37 0.12 0.08 0.14 0.20
11.85 10.74 0.10 0.11 0.20 0.25
12.03 11.51 0.11 0.12 0.22 0.28
12.22 12.02 0.08 0.16 0.26 0.32
12.41 12.63 0.09 0.15 0.28 0.33
12.60 13.57 0.09 0.18 0.33 0.39
12.78 14.90 0.10 0.19 0.36 0.42
12.97 16.55 0.10 0.21 0.39 0.45
13.16 18.42 0.09 0.21 0.40 0.46
13.34 20.43 0.09 0.24 0.46 0.53
13.53 22.56 0.11 0.26 0.51 0.58
13.71 24.80 0.12 0.28 0.55 0.62
13.89 27.20 0.12 0.31 0.60 0.69
14.07 29.78 0.14 0.33 0.66 0.75
14.25 32.62 0.15 0.36 0.71 0.81
14.43 35.74 0.17 0.40 0.81 0.92
14.61 39.30 0.19 0.44 0.85 0.97
14.79 43.35 0.19 0.49 0.95 1.08
14.96 48.06 0.22 0.54 1.06 1.21
15.14 53.50 0.21 0.59 1.15 1.31
15.31 59.48 0.25 0.65 1.27 1.45
15.48 65.73 0.25 0.70 1.36 1.55
15.66 71.74 0.30 0.73 1.42 1.63
15.82 77.12 0.32 0.79 1.58 1.79
15.99 81.53 0.30 0.92 1.78 2.03
16.16 85.05 0.31 0.99 1.93 2.20
16.32 87.75 0.32 1.09 2.06 2.35
16.65 91.18 0.34 1.24 2.28 2.62
16.81 92.06 0.47 1.23 2.33 2.68
16.96 92.23 0.42 1.28 2.47 2.82
17.27 90.85 0.40 1.28 2.50 2.84
17.58 87.90 0.38 1.40 2.67 3.04
17.87 84.20 0.31 1.43 2.64 3.02
18.15 80.26 0.42 1.47 2.61 3.02
18.29 78.33 0.43 1.54 2.65 3.09
18.43 76.55 0.42 1.42 2.61 3.00
18.70 73.70 0.36 1.32 2.51 2.86
18.95 71.67 0.45 1.28 2.47 2.81
19.20 69.68 0.42 1.27 2.38 2.73
19.44 67.06 0.34 1.26 2.34 2.68
19.67 63.70 0.32 1.31 2.30 2.66
Show more
4

Discussions

4.1
63Cu(γ, n) Reaction Cross Section

Here, we compare our measurements with available experimental [4, 6-9] and evaluated [33] data. The results are presented in Fig. 6. The uncertainty of our unfolded cross-sections is comparable to those of Fultz et al. [7] and Sund et al. [9] with monochromatic photons obtained from positron annihilation in flight. Moreover, it was significantly better than those of Owen et al. [8], Berman et al. [6] and Plaisir et al. [4] obtained with bremsstrahlung radiation. Our measurements are consistent with the data of Berman et al. [6], although the latter have only a few data points. Moreover, our data agree well with those of other groups when Sn < < 15 MeV. However, our data were visibly higher than those of other studies at > 15 MeV.

Fig. 6
(Color online) Unfolded cross section curve for 63Cu(γ, n) together the available experimental and evaluated data [6-9, 33]
pic

The integral ratio of two cross-section curves reflects their systematic differences [34]. The total cross-section integrated over the energy region of interest is defined as follows: σint=EminEmaxσ(Eγ)dEγ. (9)

We conducted experimental measurements on 197Au(γ, n) and 159Tb(γ, n) reactions at SLEGS [28]. The 197Au(γ, n) reaction data were compared with those reported by Itoh et al. [30]. The resulting σint difference was ~0.4%, suggesting the reliability of SLEGS in the measurement procedure and data analysis [28]. We calculated σint of 63Cu(γ, n) reactions for different laboratories within S1n < < 15 MeV and 15 MeV < < S2n. The results are shown in Table 2. For S1n < < 15 MeV, the relative difference between the σint value and those of the others is 4–13%, except for the data of Owen, for which the difference is 28%. In contrast, for 15 MeV < < S2n, our data are larger than the others by a factor of 0.13–0.35. Table 2 shows that the data of Owen are evidently lower than the others within the two aforementioned energy regions. Consequently, it is not a priority. Overall, our measurements are expected to clarify the inconsistency between the available experimental data for the 63Cu(γ, n) reaction.

Table 2
Integrated 63Cu(γ, n) cross section data
Author σint (mb)
  [S1n, 15 MeV] [15 MeV, S2n]
This work 83.25 369.27
Plaisir et al. [4] 86.49 277.66
Varlamov et al. [33] 93.08 320.18
Owen et al. [8] 60.21 239.32
Sund et al. [9] 76.90 318.37
Fultz et al. [7] 93.76 284.52
Show more
4.2
Radiative 62Cu(n, γ) Cross Section

γSF [35, 36] is a statistical quantity employed in the Hauser-Feshbach model of the compound nuclear reaction. The γSF in the de-excitation mode aids in determining the radiative (n, γ) cross-sections that are directly relevant to the s-process nucleosynthesis of elements heavier than iron. The downward γSF for dipole radiation at a given energy is defined as [37] fX1(Eγ)=Eγ3ΓX1(Eγ)Dl. (10) Here, X is either electric (E) or magnetic (M); ΓX1(Eγ) is the average radiation width; and Dl is the average level spacing for s-wave (l=0) or p-wave (l=1) neutron resonances.

In contrast, γSF in the excitation mode for dipole radiation [37] is defined by the average cross section for E1/M1 photoabsorption σX1(Eγ) to the final states with all possible spins and parities [36]: fX1(Eγ)=Eγ1gJ(πc)2σX1(Eγ). (11) Here, the spin factor gJ = (2J + 1)/(2J0 + 1), where J = 1 and J0 = 0 (ground state).

Above the neutron separation energy, except at energies near the neutron threshold, the total upward γSF can be determined by substituting σX1(Eγ) with the experimental (γ, n) cross-sections that dominate the photoabsorption cross-sections. According to the principle of detailed balance [38] and the generalized Brink hypothesis, the equality of the upward and downward γSF, fX1(Eγ)=fX1(Eγ)=fX1(Eγ), connect the (upward) (γ, n) cross section σγn to the (downward) γSF by [37] fX1(Eγ)=1gJπ22c2σγn(Eγ)Eγ, (12) where 1/gJπ22c2=8.674×108mb1MeV2. This relation yields the experimentally constrained γSF from the measured 63Cu(γ, n) reaction data, as indicated by the red dots in Fig. 7.

Fig. 7
(Color online) Comparison of the γSF of 63Cu calculated using the Brink-Axel Lorentzian model (blue dashed line) and SMLO model (black dashed line) for the E1 strength in TALYS with the γSF extracted from our data (red dots). γSF values (red solid line) optimized using Gnorm. The spin-flip and scissor model for the M1 strength is indicated by the pink dashed line
pic

In TALYS (version 1.96) [39, 40], various phenomenological and microscopic models have been established to describe the γSF. The Brink-Axel Lorentzian model and simple modified Lorentzian (SMLO) model [41] for the E1 strength closely approximate the experimental values. The blue and black lines in Fig. 7 represent the γSFs calculated using the aforementioned two models. The pink lines in Fig. 7 shows the spin-flip and scissor model of the M1 strength [42]. Although these two models have contributed to advancements in calculating the γSF, discrepancies between model predictions and experimental observations remain. To improve the predictive accuracy of these models, we refined the Brink-Axel Lorentzian model by incorporating the normalization factor Gnorm for γSF available in TALYS. Gnorm was optimized by minimizing χ2 and aligning the theoretical calculations of the γSF more closely with the experimental data. The expression for χ2 is given by χ2=1Ni=1N(σth,iσexp,iσerr,i)2. (13) where N represents the total number of experimental data points, and σth,i, σexp,i, and σerr,i denote the theoretical value, experimental data, and experimental error of the γSF for the i-th data point, respectively. By adjusting Gnorm, we find that the χ2 value reaches a minimum of 1.46 when Gnorm = 1.2. The red solid line in Fig. 7 represents the optimized γSF values, demonstrating closer agreement with the experimental data.

The radiative (n, γ) cross-section strongly depends on the γSF and is sensitive to the nuclear level density (NLD) model employed. We extracted the experimentally constrained γSF from our newly measured 63Cu(γ, n) reaction data and then optimized the Brink-Axel Lorentzian model for E1 strength in TALYS using Gnorm. Finally, the radiative (n, γ) cross section for 62Cu was calculated based on the Brink-Axel Lorentzian model with Gnorm optimization. The results are presented in Fig. 8 as the red band. The spin-flip and scissor model of the M1 strength was considered in the TALYS calculations. The theoretical uncertainty corresponds to the use of six NLD models [40]. A similar study was performed by Utsunomiya et al. [43], in which radiative (n, γ) cross sections of 136,137Ba isotopes were obtained. In our study, owing to the lack of experimental data on 63Cu in terms of low-lying excited levels and neutron resonance spacings, we could not effectively constrain the NLD model. Consequently, a relatively large theoretical uncertainty was obtained. To reduce the theoretical uncertainty of the (n, γ) cross sections both the γSF and NLD models should be effectively constrained. A good example can be found in Renstrom et al. [44], in which charged-particle-induced reaction data were used to constrain the NLD model.

Fig. 8
(Color online) 62Cu(n, γ) cross section calculated with TALYS code based on the Brink-Axel Lorentzian model with Gnorm optimization. The theoretical uncertainty corresponds to the use of different NLD models [39, 40]. Additional TALYS calculations based on the Brink-Axel Lorentzian model (blue line) and the SMLO model (black line) without Gnorm optimization as well as TENDL-2023 evaluations [45] (pink line) are also shown for comparison
pic

To further investigate the 62Cu(n, γ) cross section, additional TALYS calculations based on the Brink-Axel Lorentzian and SMLO models were performed without Gnorm optimization. The calculated results and available TENDL-2023 evaluations [45] are also presented in Fig. 8 for comparison, which shows a good agreement between each other. To the best of our knowledge, this is the first time that the experimentally constrained 62Cu(n, γ) cross section has been obtained. This supports the principle of detailed balance and generalized Brink hypothesis for 63Cu isotope.

5

Conclusion

We performed new measurements on the 63Cu(γ, n) cross section at energies below S2n with quasi-monochromatic and energy-tunable SLEGS γ beams. Using the unfolding iteration method, the 63Cu(γ, n) reaction data were obtained within the energy range of 11.1-19.7 MeV and the resulting uncertainty was controlled within 4%. The comparison between our measurement and previously available experimental and evaluated cross sections was discussed, helping in resolving a long-standing discrepancy between the existing 63Cu(γ, n) reaction data. Based on these new data, the experimentally constrained γSF for 63Cu was extracted, which was reasonably consistent with the TALYS calculations when considering different γSF models. Furthermore, the cross-sectional curve of the inverse reaction, 62Cu(n, γ), was obtained for the first time. Our calculations provide an alternative for extracting the (n, γ) cross sections for some unstable nuclides.

References
1.M.A. Green, C.J. Mathias, M.J. Welch, et al.,

Copper-62-labeled pyruvaldehyde bis (n4-methylthiosemicarbazonato) copper (ii): synthesis and evaluation as a positron emission tomography tracer for cerebral and myocardial perfusion

. Journal of Nuclear Medicine 31, 19891996 (1990). https://doi.org/10.1097/00004424-199012000-00013
Baidu ScholarGoogle Scholar
2.W. Luo,

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

. Nuclear Science and Techniques 27, 96 (2016). https://doi.org/10.1007/s41365-016-0094-6
Baidu ScholarGoogle Scholar
3.W. Luo, D.L. Balabanski, D. Filipescu,

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

. Nuclear Science and Techniques 27, 113 (2016). https://doi.org/10.1007/s41365-016-0111-9
Baidu ScholarGoogle Scholar
4.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

. The European Physical Journal A 48, 68 (2012). https://doi.org/10.1140/epja/i2012-12068-7
Baidu ScholarGoogle Scholar
5.Z.C. Li, Y.X. Yang, W. Luo, et al.,

Fast measurement of quasmonochromatic MeV γ-beam flux using 63Cu(γ, n)62Cu reaction at Shanghai Laser Electron Gamma Source

. Nucl. Instr. and Meth. B 559, 165595 (2025). https://doi.org/10.1016/j.nimb.2024.165595
Baidu ScholarGoogle Scholar
6.A.I. Berman, K.L. Brown,

Absolute cross section versus energy of the 63Cu(γ, n) and 63Cu(γ, 2n) reactions

. Physical Review 96, 83 (1954). https://doi.org/10.1103/PhysRev.96.83
Baidu ScholarGoogle Scholar
7.S. Fultz, R. Bramblett, J. Caldwell, et al.,

Photoneutron cross sections for natural Cu, 63Cu, and 65Cu

. Physical Review 133, B1149 (1964). https://doi.org/10.1103/PhysRev.133.B1149
Baidu ScholarGoogle Scholar
8.D. Owen, E. Muirhead, B. Spicer,

Structure in the giant resonance of 64Zn and 63Cu

. Nuclear Physics A 122, 177183 (1968). https://doi.org/10.1016/0375-9474(68)90711-2
Baidu ScholarGoogle Scholar
9.R. Sund, M. Baker, L. Kull, et al.,

Measurements of the 63Cu(γ, n) and (γ, 2n) Cross Sections

. Physical Review 176, 1366 (1968). https://doi.org/10.1103/PhysRev.176.1366
Baidu ScholarGoogle Scholar
10.M. Martins, E. Hayward, G. Lamaze, et al.,

Experimental test of the bremsstrahlung cross section

. Physical Review C 30, 1855 (1984). https://doi.org/10.1103/PhysRevC.30.1855
Baidu ScholarGoogle Scholar
11.M. Antunes, M. Martins,

Two proton and two neutron photoemission cross sections of 63Cu

. Physical Review C 52, 1484 (1995). https://doi.org/10.1103/PhysRevC.52.1484
Baidu ScholarGoogle Scholar
12.A. Varlamov, V. Varlamov, D. Rudenko, et al.,

Atlas of giant dipole resonances. Parameters and graphs of photonuclear reaction cross sections

. Tech. rep., International Atomic Energy Agency. International Nuclear Data Committee (1999)
Baidu ScholarGoogle Scholar
13.V. Varlamov, A. Davydov, M. Makarov, et al.,

Reliability of the data on the cross sections of the partial photoneutron reaction for 63,65Cu and 80Se nuclei

. Bulletin of the Russian Academy of Sciences: Physics 80, 317324 (2016). https://doi.org/10.3103/S1062873816030333
Baidu ScholarGoogle Scholar
14.A.I. Blokhin, M.B. Chadwick, T. Fukahori, et al., Handbook on Photonuclear Data for Applications. Cross Sections and Spectra. IAEA-TECDOC-1178..
15.T. Kawano, Y. Cho, P. Dimitriou, et al.,

IAEA photonuclear data library 2019

. Nuclear data sheets 163, 109162 (2020). https://doi.org/10.1016/j.nds.2019.12.002
Baidu ScholarGoogle Scholar
16.V. Varlamov, B. Ishkhanov, V. Orlin, et al.,

Evaluated cross sections of the σ(γ, nX) and σ(γ, 2nX) reactions on 112,114,116,117,118,119,120, 122,124Sn isotopes

. Bulletin of the Russian Academy of Sciences: Physics 74, 833841 (2010). https://doi.org/10.3103/S1062873810060225
Baidu ScholarGoogle Scholar
17.V. Varlamov, B. Ishkhanov, V. Orlin, et al.,

New data for the 197Au(γ, nX) and 197Au (γ, 2nX) reaction cross sections

. Bulletin of the Russian Academy of Sciences: Physics 74, 842849 (2010). https://doi.org/10.3103/S1062873810060237
Baidu ScholarGoogle Scholar
18.Z.C. Li, Y. Yang, Z.W. Cao, et al.,

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

. Nuclear Science and Techniques 34, 170 (2023). https://doi.org/10.1007/s41365-023-01330-z
Baidu ScholarGoogle Scholar
19.W. Luo, W. Xu, Q. Y. Pan, et al.,

A 4D Monte Carlo laser-Compton scattering simulation code for the characterization of the future energy-tunable SLEGS

. Nucl. Instr. and Meth. A 660, 108115 (2011). https://doi.org/10.1016/j.nima.2011.09.035
Baidu ScholarGoogle Scholar
20.Hanghua Xu, Gongtao Fan, Hailong Wu, et al.,

Interaction Chamber Design for an Energy Continuously Tunable Sub-MeV Laser-Compton Gamma-Ray Source

. IEEE Transactions on Nuclear Science 63, 906912 (2016). https://doi.org/10.1109/TNS.2015.2496256
Baidu ScholarGoogle Scholar
21.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, 5363 (2020). https://doi.org/10.11804/NuclPhysRev.37.2019043
Baidu ScholarGoogle Scholar
22.H.W. Wang, G.T. Fan, L.X. Liu, et al.,

Commissioning of laser electron gamma beamline SLEGS at SSRF

. Nuclear Science and Techniques 33, 87 (2022). https://doi.org/10.1007/s41365-022-01076-0
Baidu ScholarGoogle Scholar
23.J.H. He, Z.T. Zhao, et al.,

Shanghai Synchrotron Radiation Facility

. National Science Review 1, 171 (2014). https://doi.org/10.1093/nsr/nwt039
Baidu ScholarGoogle Scholar
24.Z.R. Hao, G.T. Fan, H.W. Wang, et al.,

A new annular collimator system of SLEGS beamline at Shanghai Light Source

. Nucl. Instrum. Methods Phys. Res. B 519, 914 (2022). https://doi.org/10.1016/j.nimb.2022.02.010
Baidu ScholarGoogle Scholar
25.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
26.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

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 1063, 169314 (2024). https://doi.org/10.1016/j.nima.2024.169314
Baidu ScholarGoogle Scholar
27.Z.R. Hao, G.T. Fan, L.X. Liu et al.,

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

. Nuclear Techniques (in Chinese) 43, 110501 (2020). https://doi.org/10.11889/j.0253-3219.2020.hjs.43.110501
Baidu ScholarGoogle Scholar
28.Z.R. Hao, L.X. Liu, Y. Zhang et al.,

Photoneutron dataset generation and analysis at SLEGS

. ChinaXiv:202412.00386 https://doi.org/10.12074/202412.00386
Baidu ScholarGoogle Scholar
29.B.L. Berman, S. Fultz,

Measurements of the giant dipole resonance with monoenergetic photons

. Reviews of Modern Physics 47, 713 (1975). https://doi.org/10.1103/RevModPhys.47.713
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

. Journal of Nuclear Science and Technology 48, 834840 (2011). https://doi.org/10.1080/18811248.2011.9711766
Baidu ScholarGoogle Scholar
31.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

. Physical Review C 90, 064616 (2014). https://doi.org/10.1103/PhysRevC.90.064616
Baidu ScholarGoogle Scholar
32.H. Utsunomiya, T. Renstrøm, G.M. Tveten, et al.,

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

. Physical Review C 98, 054619 (2018). https://doi.org/10.1103/PhysRevC.98.054619
Baidu ScholarGoogle Scholar
33.V.V. Varlamov, N.G. Efimkin, B.S. Ishkhanov, et al.,

Evaluation of Cu-64,65(γ, np) and Cu-63,65(γ, p) reaction cross sections in the energy range of giant dipole resonance and isospin splitting of the GDR of Cu nuclei

. Bull Rus. Acad. Sci. Phys 59, 911 (1995).
Baidu ScholarGoogle Scholar
34.V. Varlamov, B. Ishkhanov, V. Orlin,

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

. Physical Review C 95, 054607 (2017). https://doi.org/10.1103/PhysRevC.95.054607
Baidu ScholarGoogle Scholar
35.G. Bartholomew, E. Earle, A. Ferguson, et al.,

Gamma-ray strength functions

. Advances in Nuclear Physics 7, 229324 (1973). https://doi.org/10.1007/978-1-4615-9044-6_4
Baidu ScholarGoogle Scholar
36.J. Kristiak, E. Betàk, Neutron Induced Reactions: Proceedings of the 4th International Symposium Smolenice, Czechoslovakia, June 1985, (Springer Science & Business Media, 2012)
37.R. Capote, M. Herman, P. Obložinskỳ, et al.,

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

. Nuclear Data Sheets 110, 31073214 (2009). https://doi.org/10.1016/j.nds.2009.10.004
Baidu ScholarGoogle Scholar
38.J.M. Blatt, V.F. Weisskopf, Theortical Nuclear Physics, (John Wiley & Sons, New York, 1952)
39.A.J. Koning, D. Rochman,

Modern nuclear data evaluation with the TALYS code system

. Nuclear data sheets 113, 28412934 (2012). https://doi.org/10.1016/j.nds.2012.11.002
Baidu ScholarGoogle Scholar
40.A. Koning, S. Hilaire, S. Goriely,

TALYS: modeling of nuclear reactions

. The European Physical Journal A 59, 131 (2023). https://doi.org/10.1140/epja/s10050-023-01034-3
Baidu ScholarGoogle Scholar
41.S. Goriely, V. Plujko,

Simple empirical E1 and M1 strength functions for practical applications

. Physical Review C 99, 014303 (2019). https://doi.org/10.1103/PhysRevC.99.014303
Baidu ScholarGoogle Scholar
42.E. Balbutsev, I. Molodtsova, P. Schuck,

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

. Nuclear Physics A 872, 4268 (2011). https://doi.org/10.1016/j.nuclphysa.2011.09.013
Baidu ScholarGoogle Scholar
43.H. Utsunomiya, T. Renstrøm, G.M. Tveten, et al.,

γ-ray strength function for barium isotopes

. Physical Review C 100, 034605 (2019). https://doi.org/10.1103/PhysRevC.100.034605
Baidu ScholarGoogle Scholar
44.T. Renstrøm, H. Utsunomiya, H.T. Nyhus, et al.,

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

. Physical Review C 98, 054310 (2018). https://doi.org/10.1103/PhysRevC.98.054310
Baidu ScholarGoogle Scholar
45.

TENDL-2023 nuclear data library

., https://tendl.web.psi.ch/tendl_2023/neutron_html2024/neutron.html
Baidu ScholarGoogle Scholar
Footnote

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