logo

Resonance Analysis of 159Tb(n,γ) reaction based on the CSNS Back-n experiment

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Resonance Analysis of 159Tb(n,γ) reaction based on the CSNS Back-n experiment

De-Xin Wang
Su-Ya-La-Tu Zhang
Wei Jiang
Jie Ren
Mei-Rong Huang
Jing-Yu Tang
Xi-Chao Ruan
Hong-Wei Wang
Long-Xiang Liu
Xue Li
Dan-Dan Niu
Guo Li
Gu-Fu Meng
Yong-Shun Huang
Zhi-Long Wang
Yu Bai
Xue Yang
Nuclear Science and TechniquesVol.36, No.3Article number 43Published in print Mar 2025Available online 28 Jan 2025
4100

The neutron capture resonance parameters for 159Tb are crucial for validating nuclear models, nucleosynthesis during the neutron capture process, and nuclear technology applications. In this study, resonance analyses were performed for the neutron capture cross-sections of 159Tb measured at the China Spallation Neutron Source (CSNS) backscattering white neutron beamline (Back-n) facility. The resonance parameters were extracted from the R-Matrix code SAMMY and fitted to the experimental capture yield up to the 1.2 keV resolved resonance region (RRR). The average resonance parameters were determined by performing statistical analysis on the set of the resonance parameters in the RRR. These results were used to fit the measured average capture cross sections using the FITACS code in the unresolved resonance region from 2 keV to 1 MeV. The contributions of partial waves l=0, 1, 2 to the average capture cross-sections are reported.

Statistical analysisResonance parameters159Tb(n,γ) cross sectionγ
1

Introduction

Resonance parameters are nuclear data that play a crucial role in areas such as nuclear energy, security, and structure. These parameters are mainly obtained from neutron resonance spectroscopy and include the neutron strength function (S0), average radiation width (Γγ), and average level spacing (D0) between the resonances [1]. These parameters are critical for substantiating nuclear structure models and serve as key inputs for nuclear physics applications, such as calculating Maxwellian-averaged cross sections (MACS) at keV energies that are relevant to nuclear astrophysics [2, 3].

In stellar nucleosynthesis, the neutron capture cross sections of rare-earth isotopes constitute a fundamental component of the slow neutron capture process (s-process) and significantly influence the calculated stellar element abundances and astrophysical reaction rates [4]. For Terbium-159, a monoisotopic element, scarce neutron capture cross-sectional data are available from 1 eV to 1 MeV [5]. Additionally, the data presently accessible within this energy range are not only outdated, but also lack the requisite detail, leading to considerable inconsistencies in the resonance parameters. Previously reported experimental data shows that S0 and D0 vary from 0.9×10-4 to 1.56×10-4 and 3.21 to 4.40 eV [6-8]. Specifically, the latest evaluated nuclear data library, ENDF/B-VIII.0 [7], incorporates resonance parameters from an earlier database, JENDL-2.0, which utilized neutron capture measurements by Ohkubo et al. [9] and Mizumoto et al. [6] and data from Mughabghab [10]. Although their resonance parameters are nearly identical, JEFF-3.3 [11] and ENDF/B-VIII.0 [7] have different approaches to specific resonant structures and smooth regions.

In this study, the average resonance parameters were derived from the statistical analysis of the individual resonance parameters in the resolved resonance region (RRR) of the China Spallation Neutron Source (CSNS) backscattering white neutron beamline (Back-n) experimental neutron capture data. These results were used as input values for fitting the measured average capture cross-sections with the FITACS code from 2 keV to 1 MeV. The average capture cross-sections are given for partial waves l=0, 1, 2 and are compared with the corresponding data from the literature and the evaluation nuclear data library.

In Sect. 2, the CSNS Back-n neutron capture cross-section experiments and resonance analysis results are outlined. Section 3 presents a statistical analysis of the resolved resonance region parameters. In Sect. 4, the results are integrated with FITACS to analyze the unresolved resonance region (URR). Finally, Section 5 summarizes the discussions and conclusions.

2

Neutron Capture Measurement

2.1
Back-n facility at CSNS

The measurements were conducted at the CSNS Back-n. The CSNS comprises an 80 MeV linear accelerator, a 1.6 GeV proton synchrotron (PS), two beam transport lines, a target station, three spectrometers, and the associated instrumentation [12, 13]. At the CSNS, spallation neutrons are generated by a 1.6 GeV/c proton beam with a 41 ns full width at half maximum from the PS incident on a tungsten target. The PS operates in either double- or single-beam cluster mode, delivering pulses at 25 Hz and an average current of 64 μA. A dedicated 15 ° deflection magnet designed by CSNS steers the proton beam before the target, separating the backscattered white neutron beam from the primary proton beam [14].

The Back-n beamline is situated 80 m downstream of the CSNS proton beam. It comprises two experimental stations: 55 m (ES#1) and 76 m (ES#2). ES#2 is outfitted with two different systems for capture experiments: the 4π GTAF-II array of 40 BaF2 detectors for total absorption calorimetry and a total energy detector consisting of four low-sensitivity C6D6 detectors [15]. The neutron flux at the ES#2 sample position is 6.92×105 cm-2·s-1 over an energy range of 0.3 eV to 200 MeV distributed over three beam spot sizes: (Ф20 mm, Ф30 mm, and Ф60 mm). The neutron spectrum was characterized using a Li-Si detector and a calibrated fission chamber, exploiting the standard 6Li(n,t) and 235U(n,f) reactions, respectively [16]. Additionally, a silicon flux monitor (SiMon) comprising a thin 6LiF converter layer and eight silicon detectors positioned 20 mm upstream of the sample continuously monitored the beam intensity [17-19].

2.2
The 159Tb and Samples

In this study, neutron capture cross-sectional measurements of 159Tb were performed using a C6D6 detector array. A metallic sample of pure terbium with a diameter of SI30.0 mm and area density of 6.27×10-4 atom/barn was used for the prompt capture γ-ray measurement. A natural lead natPb with a diameter and thickness of 0.53 mm was used to evaluate the in-beam γ-ray and neutron-induced background [13]. Additionally, 197Au sample with thickness of 0.1 mm was used to normalize the neutron capture data. An empty sample holder was placed on the beam to measure the background it generates. The evaluated background spectrum was normalized using the black resonance technique, which made it possible to distinguish between background and true signals. Samples 59Co and 181Ta were employed as neutron filters. Sample preparation was performed by the Department of Nuclear Physics of the China Institute of Atomic Energy. Much More details on the experimental setup can be found in Ref. [20-26].

2.3
Background Deduction and Data Analysis

Background determination is a major challenge in experimental studies. The experimental data analysis was largely dependent on Monte Carlo simulations conducted using the Geant4 toolkit [27], which accurately reproduced the experimental configuration and formed a critical foundation for subsequent analyses. To maximize the precision of the measured gamma energy spectra, we implemented a rigorous efficiency correction that accounted for gamma energy in detection efficiency. This was achieved using the pulse-height weighting technique (PHWT) [28], which applies weighting functions to ensure that efficiency is independent of the cascade path and proportionality to the known cascade energy.

In addition to simulations, essential experimental measurements were performed using neutron filters composed of 59Co and 181Ta [20]. Here, we employed the black resonance technique to normalize the experimental background, enabling the separation of the background contributions from the true spectrum. This systematic approach is crucial for obtaining the required capture yield precision. The experimental neutron capture yield as a function of En is calculated as follows Yexp(En)=Cw(En)Cb(En)fAuΦ(En)×Ec, (1) where Cw(En) and Cb(En) are the weighted count and total background spectra, respectively, fAu is the normalization factor determined using the saturated resonance technique for the 4.9 eV resonance of the Au sample, Φ(En) is the neutron flux spectrum, and Ec is the detection efficiency of a capture event.

2.4
Resolved Resonance Region

This performs a theoretical reanalysis of 159Tb neutron capture yield from a previous experiment [13] and interprets its statistical properties. We obtained the fundamental resonance parameters by fitting the measured 159Tb(n,γ) capture yields in the resonance region using the R-matrix code SAMMY [29], based on the Reich-Moore approximation. The experimental conditions, neutron multiple scattering, self-shielding, and Doppler effects were included in the SAMMY code fitting. The resonance spin J and partial radiative and neutron widths Γn and Γγ, respectively, cannot be determined accurately using the capture measurement. Only the resonance energy and capture kernel k, which are directly related to the resonance area, can be reliably obtained from these capture measurements [30, 31]. The k defined as k=gΓnΓγΓn+Γγ, (2) where g indicates the spin (statistical) weighting factor and Γn and Γγ are the partial neutron and radiative widths, respectively. Usually, g is given by g=2J+1(2I+1)(2s+1), (3) where J is the compound nucleus resonance spin (total angular momentum) and I and s denote the target and neutron spins, respectively. In Fig. 1(a), the measured neutron capture yields at the resonance energy of 11.04 eV are fitted with the different spin J by the SAMMY code. Both J=1 and J=2 are well-fitted to the experiment, as expected. The relationships between the resonance parameters are given in the (Γ, 2gΓn0) plane at En=11.04 eV resonance for different J, as shown in Fig. 1(b). The solid red and dashed black lines indicate J=1 and J=2, respectively. The two lines of Fig. 1(b) means there are many group of values in Γn and Γγ with fitted results for one of spin J (eg. J=1 or J=2). Usually, even when both the capture and transmission data are available, accurately establishing the spin factor remains challenging [9].

Fig. 1
(Color online) (a)The SAMMY fit to the measured yield at En=11.04 eV with different spins J. (b) Resonance parameters analysis in the (Γ, 2gΓn0) plane at 11.04 eV. The red solid line and black dashed line indicate J=1 and J=2, respectively
pic

Experimental studies focused on calculating the resonance spins for 159Tb are scarce. Among the available methods, transmission measurements using polarized neutrons on polarized 159Tb nuclei have proven to be more precise in identifying the resonance J values. These measurements were reported in [32]. Given the scarcity of data and the high precision of this method, we adopted resonance spin values from [32] for our R-matrix analysis. Using this approach, we avoided some limitations of parameter extraction while relying on the most accurate spin information available.

We used SAMMY code for the resonance analysis of the 159Tb neutron capture yield data from 1 eV to 1.2 keV [13]. Each SAMMY fitting iteration provided the initial resonance parameter values, which were iteratively refined until convergence. Below 100 eV, SAMMY was initialized with the parameter set from Table VII in Ref. [13]. Above 100 eV to 1.2 keV, initial parameters were taken from the JEFF-3.3 database [11]. Figure 2 compares the 159Tb capture yield (black dots) with the SAMMY fit results (red lines). The resonance parameters are listed in Table 3.

Fig. 2
(Color online) SAMMY fits of 159Tb capture yield data from 1 eV to 1.2 keV
pic
3

Statistical Properties of the Resonance Parameters

3.1
Average Level Spacing and Average Radiative Width

The resonance distribution as a function of the neutron energy in the experimental measurements was directly related to the nuclear-level density ρ of the compound nuclei at the neutron separation energy. Specifically, ρJ for a given spin J can be derived from the number of observed resonances NJ within the neutron energy interval Δ En using ρJ=NJΔEn=1DJ. (4) where DJ represents the average level spacing of spin J. The latter can be calculated based on the cumulative distribution of the observed resonances NJ (En): NJ(En)=a+DJEn. (5) The spacing between successive resonances, which share the same total angular momentum and parity, was predicted to exhibit random behavior, which was verified experimentally. For a series of N consecutive resonances at energy Ei, the level spacing is defined as Di = Ei - Ei-1. The probability distribution of these level spacings is expected to follow the Wigner-Dyson distribution [33] as Eq.(6). p(x)dx=πx2e(πx24)dx,x=DiD. (6) The Wigner-Dyson law provides a mathematical prediction of the level-spacing distribution in chaotic systems, accurately reproducing experimental observations. For narrow-spaced resonances, the level repulsion effect leads to a distribution that disappears at small spacings, reflecting the repulsive nature of close levels. The Wigner-Dyson distribution predicts an average spacing of x=1 and variance of σ2=(4π1). Consequently, the relative uncertainty in the average-level spacing obtained from the N observed resonances can be expressed by Eq.(7): ΔDD=1N(4π1). (7) The cumulative number of experimentally observed levels as a function of neutron energy is shown in Fig. 3. The cumulative number of resonance analyses revealed differences between the data provided in Refs. [34] which is greater than 100 eV. Furthermore, when the neutron energy increased, the percentage of missed resonances increased progressively, a behavior that can be explained by statistical fluctuations. A linear fit was used to determine the average level spacing D0, which was 3.84(4) electronvolt below 100 electronvolt. The consistency of the observed spin-experimental level spacing distribution was tested by comparing this value with the predicted theoretical Wigner-Dyson distribution.

Fig. 3
(Color online) Cumulative number of levels as a function of neutron energy. The green line correspond to D0=3.78 eV recommended by Mughabghab [10]. The blue line corresponds to D0=3.21 eV, which was reported by Mizumoto et.al [6]. The red line corresponds to the adjusted fitting of currently accumulated experimental data below 100 eV, producing D0=3.84± 0.04 eV
pic

The average radiative width Γγ was calculated exclusively within the energy range below 100 eV. Figure 4 shows the radiative width obtained by analyzing all resonances obtained from the SAMMY fits. The calculated weighted average of the entire energy spectrum was Γγ=102.6(13) meV. Significant deviations of the individual values from the mean values were primarily attributed to the substantial correlations between the radiative and neutron widths. These divergences do not imply the non-statistical behavior of the radiative width because all resonances at higher energies can be satisfactorily analyzed with a fixed value corresponding to Γγ.

Fig. 4
(Color online) Fitted values for the radiative widths below 100 eV. The red-dashed line corresponds to the weighted average value Γγ=102.6(13) meV
pic
3.2
Neutron Width

The neutron width Γn of the neutron energy E0 is correlated with the average lifetime τn relative to neutron emission decay. This relationship is given by an expression reported in [35]: Γn=/τn, (8) where denotes the reduced Planck’s constant. The width can be approximately considered as the time required for the excitation energy to be concentrated on a specific neutron and the probability that a neutron penetrates the nuclear potential barrier. Thus, the neutron width can be viewed as the probability per unit time that a compound nucleus emits neutrons. Therefore, the neutron width Γn is given by Γn=p/t0, (9) where t0 is the average lifetime before the energy is concentrated on an individual neutron in a specific excited state and p is the actual potential barrier penetrability. However, the reduced neutron width Γn0 at 1 eV for s-wave neutrons can be defined as Γn0=p0/t0=Γn/E0, (10) where p0 is the number of 1 eV neutrons. Thus, insight into the properties of the nuclear lifetime distribution can be obtained through the reduced neutron width, which provides valuable information independent of energy considerations.

The neutron scattering reaction encompasses only one decay channel (ν = 1). Owing to the extreme complexity and random nature of the initial state [36], the neutron width Γn follows a zero-mean normal distribution. Therefore, the reduced neutron width Γn0 conforms to the Porter-Thomas (P-T) distribution, which can be expressed as [37] PPT(x)dx=12πcex2dx. (11) where x=Γn0/Γn0 denotes the fluctuation in the normalized reduced width. However, it is typical for comparisons between the experimental data and the expected P-T distribution to be articulated in terms of the integral of the P-T distribution, interpreted as a function of xi N(xi)=N0xiPPT(x)dx=N0(1erfxi/2). (12) where N0 is the total number of resonances.

In the present work, the distribution of neutron widths was only studied between 1 eV and 100 eV because of the lack of resonances. Figure 5 illustrates the experimentally observed and computationally derived resonance numbers with Γn0>xiΓn0 as functions of xi. Hence, our experimental data exhibit remarkably good agreement with the expected J=1-resonance behavior. This result substantiates the consistency of the resonance parameters within the energy range with the least resonance overlap.

Fig. 5
(Color online) Comparison between the experimental and expected integral distributions of the neutron widths
pic
3.3
Neutron Strength Function

In optical models [38], the neutron strength function S0, radiation width Γγ, and average level spacing D are crucial for determining high-energy neutron cross-sections. The ratio of the average level spacing to the spacing between the resonances is the neutron strength function or S0. This function is associated with the average total cross section. The neutron strength function can be expressed as shown in Eq.(13) for the s-wave neutron because of the good agreement between the experimentally reduced neutron widths and the level spacing with the theoretical P-T distribution and Wigner-Dyson distributions. S0=gΓn0D0=EE+ΔEgΓn0ΔE. (13) where g is the statistical spin factor. This implies that the neutron strength function can be calculated based on the slope of the cumulative distribution of gΓn0, which represents the neutron energy. The uncertainty of the strength function is expressed as ΔS0S0=1N(4π+1). (14) S0 is derived from a linear fit (red line) to the experimental data (black dots) in the region 1 eV - 1.2 keV, as shown in Fig. 6. Comparisons of the present S0 with the literature data are presented in Table 1. Early neutron capture experiments on 159Tb had experimental limitations, resulting in the determination of neutron strength functions ranging from 0.9×10-4 to 1.56×10-4 within a restricted energy range. The current S0 value is more reliable for applications due to improved detector performance and a wider neutron energy range in the experiment. The average resonance parameters obtained from this study and in the literature are shown in Table 2.

Fig. 6
(Color online) Cumulative sum of the reduced neutron widths as a function of neutron energy
pic
Table 1
The neutron strength function in different experiments
En(eV) Year Author & ref S0(×104)
3.34–1200 2023 This Work 1.51±0.03
Below 1200 1979 Ohkubo [9] 1.55±0.15
Below 580 1979 Ohkubo [9] 1.38±0.18
2590–3464 1978 Mizumoto [6] 1.56
21.2–580 1977 Popov [39] 1.25±0.17
21.2–753 1976 Derrien [40] 1.56±0.02
3.34–97.5 1964 Wang [8] 0.90±0.30
3.34–156 1955 Harvey [35] 1.50±0.20
Show more
Table 2
Average resonance parameters obtained in this study and in the literature
    D0 (eV) S0(× 10-4) Γγ (meV)
This Work 2023 3.84(4) 1.51(3) 102.6(13)
Muhabghab 2018 [34] 3.84(16) 1.55 (15) 101(2)
Ohkubo 1978 [9] 4.4 1.55(15) 107(7)
Mizumoto 1978 [6] 3.21(1) 1.56 97
ENDF/B-VIII.0 2018 [7] 3.21 1.207 97(7.5)
Show more
4

Average Resonance Cross Section

In the URR, the resonances start to overlap as the excitation energy of the compound nuclei increases. While the resonance structures are still visible, their overlap becomes too large to allow for a good match because the intrinsic widths are close to the distance between the neutron cross-sections of neighboring resonances.

Therefore, analyzing average cross sections in the URR is more appropriate than analyzing indirect quantities like capture or transmission yields. The relationship between the average neutron capture yield Y(En), average cross-section σγ(En), and target thickness n can be represented by the following equation: Y(En)=f(En)nσγ(En). (15) where f(En) is the correction factor for the multiple scattering and self-shielding of a target of thickness n.

Currently, SAMMY cannot implement multiple scattering corrections in the URR. Self-shielding and multiple scattering effects are accounted for by f(En) factors, which are computed using the SESH code, which is a Monte Carlo code that takes nuclear properties, resonance parameters, and sample geometry specifications as inputs [41]. It then uses these values to generate an average cross-section via the Hauser-Feshbach equation and subsequently calculates the average number of collisions before incident neutron capture (or escape). After fitting the SESH data using the function a/Eb, the magnitude of the correction was found to be less than the expected 1% over the entire energy range, as shown in Fig. 7. Therefore, after subtracting the point-by-point background using Eq. (15), the average capture cross-section was calculated directly from the capture yield. Because the normalization of capture yields was shown to be accurate and consistent in RRR, the same normalization was applied in URR.

Fig. 7
(Color online) Correction factor for multiple scattering and self-shielding calculated with the SESH code
pic

The SAMMY code incorporates Fröhne’s FITACS with minor modifications [42]. FITACS uses the Hauser-Feshbach theory with width fluctuations, each partial wave has adjustable parameters for the neutron strength function Sl, the level spacing D, the average radiation width Γγ, and the long-range level parameter R from the potential scattering radius R0, where R0=a(1R) and a is the nuclear radius given by a=1.23 A1/3+0.8 fm. These quantities can easily be determined by fitting the experimental neutron capture yield to the initial values determined in RRR. In this study, the FITACS code was used to analyze the contributions of different partial waves to the average cross sections. In FITACS, the level-information data for commonly used inelastic channels are obtained from the ENSDF [43] or JEFF-3.3 [11] databases. The contributions of the first three partial waves were sufficient up to 1 MeV. In the fitting process, the initial values of the resonance parameters were adopted from the present results of the RRR analysis of the s-wave and the data of Mughabghab [34] for the p- and d-waves, respectively.

The FITACS-calculated average cross-section and contributions from the first three partial waves (l=0, 1, 2) are shown in Fig. 8. The fitted values obtained from the FITACS code were S0=1.51(3)× 10-4, S1=1.83(5)× 10-4, and S2=1.55(21)× 10-4 [13]. A comparison of various average resonance parameters used to describe URR is presented in Table 3. The calculations indicate that the cross section below 50 keV is dominated by the s-wave. Between 50 and 300 keV, the s-wave and p-wave contributions are comparable, while the d-wave contribution starts to appear above 100 keV. The importance of the p-wave effect is evident at energies as low as a few keV. The quenching above 50 keV is likely due to competition from inelastic scattering processes, which become significant at higher energies.

Fig. 8
(Color online) The measured 159Tb capture cross sections and the FITACS code calculation results. The black dots are the measured results from this work; the red solid line is the calculation result from the FITACS code. The green, blue, and pink lines represent the relative contributions of the s-wave, p-wave, and d-wave neutrons (l = 0, 1, 2), respectively
pic
Table 3
Resonance parameters up to 100 eV neutron energy for 159Tb(n, γ) reaction
ER (eV) J l This work JEFF-3.3
      Γn (meV) Γγ (meV) Γn (meV) Γγ (meV)
3.34 ± 0.03 2 0 0.34 ± 0.02 102.12 ± 0.06 0.34 103.00
4.97 ± 0.13 1 0 0.07 ± 0.05 100.58 ± 0.04 0.08 103.00
11.04 ± 0.03 2 0 4.97 ± 0.15 105.45 ± 0.02 7.69 99.00
14.40 ± 0.30 2 0 0.12 ± 0.04 98.36 ± 0.08 0.19 105.00
21.19 ± 0.13 1 0 1.23 ± 0.15 103.15 ± 0.23 1.14 102.00
24.53 ± 0.07 2 0 3.77 ± 0.16 128.35 ± 1.40 5.32 116.00
27.55 ± 0.34 2 0 0.48 ± 0.09 95.68 ± 0.85 0.83 102.00
33.83 ± 0.12 1 0 3.24 ± 0.52 102.32 ± 0.48 2.61 98.00
40.81 ± 0.63 1 0 0.55 ± 0.18 105.78 ± 1.73 0.84 101.00
43.69 ± 0.13 2 0 4.33 ± 0.28 103.54 ± 1.25 5.90 97.00
46.04 ± 0.14 2 0 7.62 ± 0.52 113.48 ± 1.29 13.94 109.00
50.16 ± 0.24 2 0 1.80 ± 0.12 102.86 ± 0.61 1.91 96.00
51.61 ± 0.38 1 0 1.14 ± 0.21 98.36 ± 1.43 0.84 96.00
54.08 ± 0.53 2 0 0.41 ± 0.16 85.51 ± 1.23 0.83 78.00
57.45 ± 0.36 1 0 2.09 ± 0.38 102.48 ± 1.00 2.20 99.00
58.77 ± 0.29 2 0 1.38 ± 0.22 98.65 ± 0.79 1.59 96.00
65.17 ± 0.18 2 0 7.90 ± 0.54 98.25 ± 1.44 12.61 96.00
66.56 ± 0.39 1 0 1.82 ± 0.51 93.22 ± 1.10 3.52 98.00
73.83 ± 0.14 2 0 14.31 ± 0.82 102.34 ± 1.28 19.09 98.00
76.45 ± 0.29 1 0 7.35 ± 0.98 116.67 ± 2.79 6.92 108.00
77.98 ± 0.37 2 0 4.85 ± 0.92 101.13 ± 2.56 7.25 96.00
78.80 ± 0.67 2 0 1.06 ± 0.62 95.15 ± 2.00 2.68 85.00
88.44 ± 0.42 2 0 3.54 ± 0.66 86.62 ± 1.11 3.37 70.00
90.60 ± 0.33 2 0 11.72 ± 1.75 105.68 ± 2.74 6.84 90.00
97.24 ± 0.31 1 0 24.37 ± 5.12 118.76 ± 4.77 38.21 101.00
Show more
5

summary and Conclusion

The neutron capture yield for 159Tb was measured using at the CSNS Back-n facility with an energy detection system (C6D6) from 1 electronvolt to 1 M electron volt. The resonance parameters for 159Tb were analyzed using the multilevel R-matrix Bayesian code SAMMY between 1 eV and 1.2 keV and used for statistical analysis to determine the average quantities required for the cross-section model calculations. The average reduced neutron widths Γn0 and average level spacing D0 for 159Tb were well-fitted by the theoretical Porter-Thomas and Wigner distributions, respectively. The average radiation width Γγ was calculated to be 102.6(13) meV. The s-wave neutron strength function S0 derived from Γn0 and D0 was 1.51(3)× 10-4. Statistical analysis of the average capture cross section in the URR was performed using the FITACS code as the input of the RRR average resonance parameters. The results show that the dominant contribution to the cross-section comes from the s-wave over the entire energy range. The contributions of the s- and p-waves should also be considered above 50 and 300 keV, respectively. The present work may provide guidance for the reevaluation of the updated nuclear data library and the improvement of nuclear theoretical models.

References
1. S. I. Sukhoruchkin and Z. N. Soroko, in Neutron Resonance Parameters, ed. by H. Schopper(Springer, Berlin Heidelberg New York, 2015) Supplement to Volume I/24, p. 2181. https://doi.org/10.1007/978-3-662-45603-3
2. Z. Y. Bao, F. Käppeler,

Neutron capture cross sections for s-process studies

. At. Data Nucl. Data Tables 36411 (1987). https://doi.org/10.1016/0092-640X(87)90011-8
Baidu ScholarGoogle Scholar
3. G. Noguere, O. Bouland, J. Heyse et al.,

s-wave average neutron resonance parameters of 175Lu+n

. Phys. Rev. C 100, 065806 (2019).https://doi.org/10.1103/PhysRevC.100.065806
Baidu ScholarGoogle Scholar
4. K. K. Seth,

S-wave neutron strength functions

. Nucl. Data Sheets. Sections A 2, 299 (1966). https://doi.org/10.1016/S0550-306X(66)80008-3
Baidu ScholarGoogle Scholar
5. V. V. Zerkin, and B. Pritychenko,

The experimental nuclear reaction data (EXFOR): Extended computer database and Web retrieval system

. Nucl. Instrum. Methods. A 888, 31 (2018).https://doi.org/10.1016/j.nima.2018.01.045
Baidu ScholarGoogle Scholar
6. M. Mizumoto, R.L. Macklin, J. Halperin,

Neutron capture cross section of 159Tb from 2.6 to 700 keV

. Phys. Rev.C 17, 522 (1978) https://doi.org/10.1103/PhysRevC.17.522
Baidu ScholarGoogle Scholar
7. D. A. Brown, M. B. Chadwick, R. Capote et al.,

ENDF/B-VIII. 0: the 8th major release of the nuclear reaction data library with CIELO-project cross sections, new standards and thermal scattering data

. Nucl. Data Sheets 148, 1 (2018). https://doi.org/10.1016/j.nds.2018.02.001
Baidu ScholarGoogle Scholar
8. N. Y. Wang, N. Iliescu, E.N. Karzhavina et al.

Neutron resonances in praseodymium and terbium

. Soviet Physics J. Exptl. Theoret. Phys. (U.S.S.R.) 47, 43 (1964).
Baidu ScholarGoogle Scholar
9. M. Ohkubo, Y. Kawarasaki,

Slow neutron resonances in terbium-159

. J. Nucl. Sci. Technol. 16, 10 (1979). https://doi.org/10.1080/18811248.1979.9730969
Baidu ScholarGoogle Scholar
10. S.F. Mughabghab, in Atlas of Nntron Cross Section, Vol.I, Part B, 3th edn. (Academic Press, New York, 1984), pp.121-128
11. A.J.M. Plompen, O. Cabellos, C. De Saint Jean et al.,

The joint evaluated the fission and fusion nuclear data library JEFF-3.3

. Eur. Phys. J. A 56, 181 (2020). https://doi.org/10.1140/epja/s10050-020-00141-9
Baidu ScholarGoogle Scholar
12. J. Y. Tang, Q. An, J. B. Bai et al.,

a back-n white-neutron source at CSNS and its applications

. Nucl. Sci. Tech. 32, 1 (2021). https://doi.org/10.1007/s41365-021-00846-6
Baidu ScholarGoogle Scholar
13. S. Zhang, G. Li, W. Jiang et al.,

Measurement of the 159Tb (n,γ) cross-section at the CSNS Back-n facility

. Phys. Rev. C 107, 045809 (2023). https://doi.org/10.1103/PhysRevC.107.045809
Baidu ScholarGoogle Scholar
14. J.Y. Tang, R. Liu, G.H. Zhang et al.,

Initial years’ neutron-induced cross-section measurements at the CSNS Back-n white neutron source

. Chin. Phys. C 45, 062001 (2021). https://doi.org/10.1088/1674-1137/abf138
Baidu ScholarGoogle Scholar
15. J. Ren, X. C. Ruan, W. Jiang et al.,

Background study of (n, γ) cross-sectional measurements with C6D6 detectors at CSNS Back-n

. Nucl. Instr. and Meth. A 985, 164703 (2021). https://doi.org/10.1016/j.nima.2020.164703
Baidu ScholarGoogle Scholar
16. Y.H. Chen, G.Y. Luan, J. Bao et al.

Neutron energy spectrum measurement of back-n white neutron source at CSNS

. Eur. Phys. J. A 55, 1 (2019). https://doi.org/10.1140/epja/i2019-12808-1
Baidu ScholarGoogle Scholar
17. X. X. Li, L. X. Liu, W. Jiang et al.,

New experimental measurement of natEr(n, γ) cross-sections between 1 and 100 eV

. Phys. Rev.C 104, 054302 (2021) https://doi.org/10.1103/PhysRevC.104.054302
Baidu ScholarGoogle Scholar
18. X.R. Hu, G.T. Fan, W. Jiang et al.,

Measurements ofthe 197Au(n, γ) cross-section up to 100keV at the CSNS Back-nfacility

. Nucl. Sci. Tech. 32, 1 (2021). https://doi.org/10.1007/s41365-021-00931-w
Baidu ScholarGoogle Scholar
19. X. X. Li, L. X. Liu, W. Jiang et al.,

The experimental determination of the neutron resonance peak of 162Er at 67.8 eV

. Phys. Rev.C 106, 065804 (2022) https://doi.org/10.1103/PhysRevC.106.065804
Baidu ScholarGoogle Scholar
20. D. X. Wang, S. Zhang, W. Jiang et al.,

neutron capture cross-sectional measurements for natLu with different thicknesses

. Acta. Phys. Sin. 71, 072901 (2022). https://doi.org/10.7498/aps.71.20212051
Baidu ScholarGoogle Scholar
21. J. Ren, X.C. Ruan, W. Jiang et al.,

the neutron capture cross-section of 169Tm measured at the CSNS Back-n facility in the energy region from 30 to 300 keV

. Chin. Phys. C 46, 044002 (2022). https://doi.org/10.1088/1674-1137/ac4589
Baidu ScholarGoogle Scholar
22. G.L. Yang, Z.D. An, W. Jiang et al.,

Measurement of Br(n,γ) cross-sections up to stellar s-process temperatures at CSNS Back-n

. Nucl. Sci. Tech. 34, 180 (2023). https://doi.org/10.1007/s41365-023-01337-6
Baidu ScholarGoogle Scholar
23. G.L. Yang, Z.D. An, W. Jiang et al.,

Measurement of neutron capture cross sections of rhenium up to stellar s-and r-process temperatures at the China Spallation Neutron Source Back-n facility

. Phys. Research 6, 013225 (2024). https://doi.org/10.1103/PhysRevResearch.6.013225
Baidu ScholarGoogle Scholar
24. J. M. Xue, S. Feng, Y. H. Chen et al.,

Measurement of the neutron-induced total cross sections of natPb from 0.3 eV to 20 MeV on the Back-n at CSNS

. Nucl. Sci. Tech. 35, 18 (2024).https://doi.org/10.1007/s41365-024-01370-z
Baidu ScholarGoogle Scholar
25. J.C. Wang, J. Ren, W. Jiang et al.,

The in-beam gamma rays of CSNS Back-n are characterized by a black resonance filter

. Nucl. Sci. Tech. 35, 164 (2024).https://doi.org/10.1007/s41365-024-01553-8
Baidu ScholarGoogle Scholar
26. X. K. L, Z. D. An, W. Jiang et al.,

Measurement of the 141Pr(n,γ) cross section up to stellar s-process temperatures at the China Spallation Neutron Source Back-n facility

, Phys. Rev.C 108, 035802 (2023). https://doi.org/10.1103/PhysRevC.108.035802
Baidu ScholarGoogle Scholar
27. S. Agostinelli, J. Allison, K. Amako et al.,

GEANT4–a simulation toolkit

. Nucl. Instr. and Meth. A. 506, 250 (2003). https://doi.org/10.1016/S0168-9002(03)01368-8
Baidu ScholarGoogle Scholar
28. R. L. Macklin, J. H. Gibbons,

Capture cross-sectional studies for 30—220-keV neutrons using a new technique

. Physical Rev. 159, 4 (1967). https://doi.org/10.1103/PhysRev.159.1007
Baidu ScholarGoogle Scholar
29. N. M. Larson,

Updated users guide for SAMMY: Multilevel The R-matrix fits the neutron data using Bayes equations. No. ORNL/TM- 2809179/R8

(Oak Ridge Naional Laboratory 2008)
Baidu ScholarGoogle Scholar
30. S. Amaducci, N. Colonna, L. Cosentino et al.,

Measurement of the 140Ce(n, γ) cross section at nTOF and its astrophysical implications for the chemical evolution of the universe

. Phys. Rev. Lett. 132, 122701(2024). https://doi.org/10.1103/PhysRevLett.132.122701
Baidu ScholarGoogle Scholar
31. C. Lederer-Woods, J. Andrzejewski, U. Battino et al.,

Measurement of the 70Ge(n, γ) cross-section up to 300 keV at the CERN nTOF facility

. Phys. Rev. C 100, 045804 (2019).https://doi.org/10.1103/PhysRevC.100.045804
Baidu ScholarGoogle Scholar
32. V. P. Alfimenkov, S.B. Borzakov, J. Wierzbicki et al.,

Investigation of the spin dependence of neutron cross sections and strength functions for rare earth nuclei in experiments with polarized neutrons and nuclei

. Nucl. Phys. A 376, 2 (1982). https://doi.org/10.1016/0375-9474(82)90062-87
Baidu ScholarGoogle Scholar
33. E.P. Wigner,

Conference on Neutron Physics by Time of Flight (Gatlinburg, Tennessee, November 1956), Report ORNL-2309

(Oak Ridge Naional Laboratory, 1957)
Baidu ScholarGoogle Scholar
34. S. F. Mughabghab, Atlas of Neutron Resonances, 6th edn. (Elsevier, Amsterdam, 2018), pp.178https://doi.org/10.1016/B978-0-44-463780-2.00015-3
35. J.A. Harvey, D.J. Hughes, R.S. Carter et al.

Spacing and neutron widths of nuclear energy levels

. Phys. Rev. 99, 10 (1955). https://doi.org/10.1103/PhysRev.99.10
Baidu ScholarGoogle Scholar
36. C.E. Porter and N. Rosenzweig,

Repulsion of energy levels in complex atomic spectra

. Phys. Rev. 120, 1698 (1960). https://doi.org/10.1103/PhysRev.120.1698
Baidu ScholarGoogle Scholar
37. C. E. Porter and R. G. Thomas,

Fluctuations in the nuclear reaction widths

. Phys. Rev. 104, 483 (1956). https://doi.org/10.1103/PhysRev.104.483
Baidu ScholarGoogle Scholar
38. L.C. Gomes,

Imaginary part of the optical potential

. Phys. Rev. 116, 1226 (1959). https://doi.org/10.1103/PhysRev.116.1226
Baidu ScholarGoogle Scholar
39. A.B. Popov, H. Faikov, et al., Yad.Fiz 26, 14 (1977).
40. J. Blons, H. Derrien.

Analyse multiniveaux des sections efficaces totale et de fission de 241Pu de 1 à 104 eV

. Journal de Physique 37, 6 (1976). https://doi.org/10.1051/jphys:01976003706065900
Baidu ScholarGoogle Scholar
41. F. H. Frohner.,

SESH Computer Code. Gulf General Atomic Report, GA-8330

(1968).
Baidu ScholarGoogle Scholar
42. F. H. Frohner., FITACS computer code. Rep. ANL-83-4 (1983).
43. M. R. Bha. Nuclear Data for Science and Technology, ed. by s. m. Qaim, (Springer, Berlin Heidelberg New York, 1992), p.1 https://doi.org/10.1007/978-3-642-58113-7-227
Footnote

The authors declare that they have no competing interests.