Skip to main content

ORIGINAL RESEARCH article

Front. Astron. Space Sci., 08 November 2023
Sec. Nuclear Physics​
This article is part of the Research Topic Cross Section Data of Interest for Nuclear Astrophysics: Experimental and Theoretical Status, and Perspectives View all 6 articles

Study of the isomeric yield ratio in the photoneutron reaction of natural holmium induced by laser-accelerated electron beams

Jingli Zhang&#x;Jingli Zhang1Wei Qi&#x;Wei Qi2Wenru FanWenru Fan1Zongwei CaoZongwei Cao1Kaijun LuoKaijun Luo1Changxiang TanChangxiang Tan1Xiaohui ZhangXiaohui Zhang2Zhigang DengZhigang Deng2Zhimeng ZhangZhimeng Zhang2Xinxiang LiXinxiang Li1Yun YuanYun Yuan1Wen Luo
Wen Luo1*Weimin Zhou
Weimin Zhou2*
  • 1School of Nuclear Science and Technology, University of South China, Hengyang, China
  • 2Science and Technology on Plasma Physics Laboratory, Laser Fusion Research Center, China Academy of Engineering Physics, Mianyang, China

Introduction: An accurate knowledge of the isomeric yield ratio (IR) induced by the photonuclear reaction is crucial to study the nuclear structure and reaction mechanisms. 165Ho is a good candidate for the investigation of the IR since the Ho target has a natural abundance of 100% and the residual nuclide has a good decay property.

Methods: In this study, the photoneutron production of 164m, gHo induced by laser-accelerated electron beams is investigated experimentally. The γ-ray spectra of activated Ho foils are off-line detected. Since the direct transitions from the 164mHo are not successfully observed, we propose to extract the IRs of the 164m, gHo using only the photopeak counts from the ground-state decay.

Results: The production yields of 164m, gHo are extracted to be (0.45 ± 0.10) × 106 and (1.48 ± 0.14) × 106 per laser shot, respectively. The resulting IR is obtained to be 0.30 ± 0.08 at the effective γ-ray energy of 12.65 MeV.

Discussion: The present data, available experimental data, and TALYS calculations are then compared to examine the role of the excitation energy. It is found that besides the giant dipole resonance, the excitation energy effect also plays a key role in the determination of the IRs.

1 Introduction

Nuclear isomers have been widely studied due to their fascinating applications, including medical imaging (Habs et al., 2011; Pan et al., 2021), nuclear clocks (Peik et al., 2021), and nuclear batteries (Prelas et al., 2014). They still play a crucial role in various aspects of astrophysical nuclear reactions (Hayakawa et al., 2008; Zilges et al., 2022). In many nuclear reactions, the residual nuclei have isomeric states with narrow energy levels and relatively long half-lives. The isomeric cross-section or yield ratio (IR) of high-spin to low-spin states of the residual nucleus provides valuable information about nuclear structure and reaction mechanisms, such as the transfer of angular momentum, the spin dependence of nuclear level density, the amelioration in γ-ray transition theory, and tests of different models (Naik et al., 2016; Rahman et al., 2020). In addition, the IR plays a key role in calculating the total production cross-section of the residual products when the production cross-section of one isomer is known in advance.

The excitation energy and angular momentum of incident particles can significantly affect the IR values of the residual products. The IRs have been studied in nuclear reactions induced by different incident particles, such as photon (Rahman et al., 2016), proton (Hilgers et al., 2007), neutron (Luo et al., 2014), and alpha (Kim et al., 2015). Compared to other particles, photons carry a smaller angular momentum of 1ℏ or 2ℏ. Furthermore, intense bremsstrahlung photons can be readily produced using radio-frequency (RF) electron accelerators. As a result, the photonuclear reactions seem to be a good tool to investigate the effect of the excitation energy on the IR. Kolev et al. (1995) deduced the experimental IRs of (γ, 3n)110m, gIn, (γ, n)164m, gHo, and (γ, 3n) 162m, gHo by a bremsstrahlung source with an end-point energy of 43 MeV. Thiep et al. (2011) determined the IRs of 165Ho(γ, n)164m, gHo and 175Lu (γ, n)174m, gLu reactions in the bremsstrahlung energy region from 14 to 25 MeV. Do et al. (2013) measured the IRs of 164m, gHo and 162m, gHo via 165Ho(γ, n) and 165Ho(γ, 3n) reactions in the bremsstrahlung energy region from 45 to 65 MeV. It is noticeable that the available IRs for the 165Ho(γ, n)164m, gHo reactions are still scarce. It particularly lacks experimental data in the energy region below 11 MeV. With the rapid development of high-intensity laser technology (Danson et al., 2019), laser–plasma interactions are used to study various nuclear phenomena (Schlenvoigt et al., 2008; Günther et al., 2022; Cao et al., 2023). Recently, the efficient production of nuclear isomers, including 113m, 115mIn and 93mMo, has been studied experimentally using the laser-accelerated electron beam (e beam) (Feng et al., 2022; Fan et al., 2023 under review).

In this study, we experimentally investigate the production of 164m, gHo by laser-induced photoneutron reactions. The γ-ray spectra of the activated Ho foils are detected by an offline γ-ray spectrometry technique. Since the direct transitions from 164mHo were not successfully observed, we propose to extract the ground and isomeric yields of 164Ho using only the photopeak counts from the ground-state decay. We should note that this approach differs from determining the counts of two photopeaks that directly characterize the isomeric and ground states. The IR value of 165Ho(γ, n)164m, gHo is obtained for a given excitation energy. The present and similar literature data on IRs are compared to examine the role of excitation energy. Furthermore, the cross-section and IR curves of the 165Ho(γ, n)164m, gHo reaction are calculated by the TALYS software to examine the compatibility of the theoretical model with the experimental data.

2 Experimental setup

The 164m, gHo production experiment was performed on the XingGuang-III laser facility at the Laser Fusion Research Center in Mianyang. The experimental setup is schematically shown in Figure 1A. An intense laser pulse with a duration of ∼0.8 ps and energy of ∼100 J was focused by an f/2.6 off-axis parabola (OAP) mirror (Robbie et al., 2018) onto a supersonic gas jet with a well-defined uniform density distribution (Feng et al., 2022). In the first stage of the experiment, high-charge multi-MeV e beams were produced during the laser–gas interactions. An image plate (IP) stack with a central hole was used to measure the spatial distribution of the laser-accelerated e beam. It should be noted that the IP stack is composed of seven IPs, with each IP being stuck on a tantalum foil with a thickness of 0.5 mm. Meanwhile, an electron magnetic spectrometer (EMS) was placed downstream of the IP stack to accurately diagnose the energy of the e beam passing through the central hole of the IP stack. In the second stage of the experiment, a metal stack composed of Ta foil and stacked Ho foils was installed, and both the IP stack and EMS were uninstalled. The Ta foil is 2 mm thick, in which energetic bremsstrahlung photons are generated. The stacked Ho foils used for activation have 10 layers in total, with each layer having a thickness of 1 mm and a natural abundance of 99.99% (Inagaki et al., 2020). During the activation, the bremsstrahlung radiations irradiate the Ho foils, successfully triggering photoneutron reactions and then producing a large number of 164m, gHo, as shown in Figure 1B. After the activation, the Ho foils are taken out from the target chamber of the XingGuang-III laser facility. The activation spectra are recorded using a high-purity germanium (HPGe) detector. The response of the HPGe detector has been well calibrated by standard γ-ray sources, including 60Co, 152Eu, 133Ba, 226Ra, 137Cs, and 241Am. In order to reduce the self-absorption effect induced by the stacked Ho foils, the 10 layers are spread out at the surface of the Al window of the HPGe detector.

FIGURE 1
www.frontiersin.org

FIGURE 1. Experimental setup for the production of 164m, gHo at the XingGuang-III laser facility (not to scale) (A). Schematic view of the photoneutron production of 164m, gHo (B) and the partial-level scheme of 164m, gHo and the decay property of 164gHo (not to scale) (C). As the laser-accelerated e beam fires to the metal stack (Ta + Ho), a large number of bremsstrahlung photons are generated, and subsequently, the Ho stacks are activated via photoneutron reactions, producing 164Ho in both the ground state (Jπ = 1+) and isomeric state (Jπ = 6) (Singh et al., 2018). The 6 isomer in 164Ho decays, via only internal decay, to the ground state. The decay from the 6 isomer to the 3+ excited state is an E3 transition. This 3+ excited state subsequently de-excites to the ground state. The resulting transition energies are 94.0, 56.6, and 37.3 keV. Finally, the ground-state decays to the daughter nucleus 164Dy or 164Er, emitting two characteristic γ-rays at 73.4 and 91.4 keV.

3 Isomeric yield ratio determination

The temporal evolution of the numbers of nuclei that are formed in the isomeric and ground states is described by the following kinetic equations (Thiep et al., 2011):

dNmdt=pmλmNm,dNgdt=pg+ηλmNmλgNg,(1)

where the subscripts m and g designate the isomeric and ground states of 164Ho, respectively; pm and pg are the production rates leading to 164Ho in the isomeric and ground states; Nm and Ng are the numbers of nuclei in the corresponding states; λm and λg are the constants of nuclear decay; and η is the transition coefficient from the isomeric state to the ground state. These equations describe the processes involved in the direct production of the desired isotopes during the target irradiation and the reduction of the number of nuclei as a result of their radioactive decay. The production rate px with x being the subscript m or g (similarly hereinafter) can be written as (Kolev et al., 1995)

px=N0EthEmaxφEγσxEγdEγ=N0Φσx,(2)

where N0 is the number of target nuclei, φEγ represents the bremsstrahlung photon flux; σxEγ is the reaction cross-section leading to the formation of 164Ho in both the isomeric and ground states; Eth and Emax are the reaction threshold and bremsstrahlung end-point energy, respectively; Φ=EthEmaxφEγdEγ is the integrated photon flux; and σx=EthEmaxφEγσxEγdEγ/EthEmaxφEγdEγ is the flux-weighted average cross-section leading to the isomeric or ground state.

Since the gamma spectroscopy method is used in the experiment, the photopeak counts (Cx) of the characteristic γ-ray of interest can be readily obtained over the detection time td. When taking into account the irradiation time tirr and the cooling time tc, the solution of Eq. 1 in three time intervals (tirr, tc, and td) and the consequent integration of the relevant activity over the td lead to

Cm=ImεmApm,(3a)
Cg=IgεgBpg+Dpm,(3b)

where Ix and εx are the branching intensity and source-peak detection efficiency of the characteristic γ-rays to be detected, respectively. The other variables are listed as follows: A=1λm1eλmtirreλmtc1eλmtd, B=1λg1eλgtirreλgtc1eλgtd, and D=ηλgλmλgλm1eλmtirreλmtc1eλmtdλmλg1eλgtirreλgtc1eλgtd.

In the case of the bremsstrahlung photon, the expression of the IR reads (Jonsson et al., 1977) as follows:

IR=σmσg=pmpg.(4)

Usually, the IR in a nuclear reaction is determined by measuring the counts of photopeaks that characterize the isomeric and ground states, respectively. When one or more photopeaks induced by the isomeric state are directly detected, the px and the resulting IR values can be readily obtained by solving Eq. 3 and Eq. 4, respectively (Rahman et al., 2020). In the case of the Ho sample, the photopeak at 37.3 keV is popularly used to characterize the isomeric state, while that at 73.4 and 91.4 keV is used to characterize the ground state, as shown in Figure 1C. In our experiment, the 37.3 keV photopeak was not successfully observed. This is because only single-shot irradiation is performed, and both the σm and εm values are relatively small. As a result, the abovementioned approach used to extract the IR value becomes invalid. It is shown in Eq. 3b that both the isomeric and ground states contribute to the Cg value, which varies with the td. It suggests that the px values can also be obtained by solving only Eq. 3b at two different time instants. More specifically, the px values can be deduced by solving the simultaneous equations of Eq. 3b at tdi and tdj, with i and j denoting two arbitrary time instants.

pg=DjYiDiYjBiDjBjDi,pm=BiYjBjYiBiDjBjDi,(5)

where Yi=Cgi/Igεg. This indicates that using only the photopeak counts from the ground-state decay, the IR value can be obtained. Furthermore, a group of data on IR can be obtained by reasonably changing the time increments. According to the error propagation, the uncertainties of the px values can be determined by the following formula:

σpg=Dj2σYi2+Di2σYj2BiDjBjDi2,σpm=Bi2σYj2+Bj2σYi2BiDjBjDi2.(6)

As a result, the uncertainty of the IR value can be written as

σIR=IRσpg2pg2+σpm2pm2.(7)

4 Results and discussion

4.1 Electron spectra

In our experiment, high-energy electrons are mainly produced by the parametrically enhanced direct laser acceleration (Cao et al., 2023). As the electron yield and charge are sensitive to the plasma density, the ebeam generation can be optimized by adjusting the backing pressure of the gas jet. Figure 2A shows the energy distributions of truncated ebeams recorded by the EMS at two backing pressures of 2.0 and 2.6 MPa. The spectral pattern of the ebeam can be described by a Boltzmann distribution dNdEeETe, where the Te is the slope temperature (Qi et al., 2019). The fitting results show that the Te values are 7.8 MeV and 4.8 MeV for backing pressures of 2.0 MPa and 2.6 MPa, respectively. In addition, the charge of the laser-accelerated electrons higher than 1 MeV is Qe ∼ 42 nC at 2.0 MPa, which is 2.5 times higher than that at 2.6 MPa. The spatial distribution of the ebeam is shown in Figure 2B. It is visibly seen that a bright spot is located beneath the central hole. According to the experimental arrangement and the spot size of the e beam, the laser-accelerated ebeam has an angular divergence of approximately 200 mrad (FWHM).

FIGURE 2
www.frontiersin.org

FIGURE 2. Spectral distributions of the laser-accelerated e beam diagnosed by the EMS (A), spatial distribution of the laser-accelerated e beam recorded at the backing pressure of 2.0 MPa (B), and the electron charge in dependence on the electron energy (C).

Generally, only high-energy electrons can be used to induce isomer production. However, the bright electron spot is located beneath the central hole, as shown in Figure 2B. This suggests that most of the electrons with high energy did not pass through the central hole of the IP stack and were not recorded by the EMS in our experiment. To understand more about the spectral pattern of the laser-accelerated e beam, we utilized the Geant4 toolkit (Agostinelli et al., 2003) to simulate the attenuation of monoenergetic electrons inside the IP stack. For a given energy, when the number of incident electrons reduces by a factor of 0.9, such energy is regarded as the minimum energy recorded by each IP. Then, the spectral distribution of the e beam can be figured out but with a relatively large uncertainty. Similar studies have been conducted by Bonnet et al. (2013) and Nishiuchi et al. (2020). Figure 2C shows the simulated electron spectral distribution, which matches well with the Boltzmann distribution. Note that the uncertainty represents the detectable energy range for each IP. The slope temperature is fitted to be 16.2 MeV, which is two times higher than the one recorded by the EMS. This is because the bright e beam was not centered with the hole on the IP stack so that the EMS only detected the low-energy part of the e beam, as presented earlier. Since the slope temperature is sufficiently high, such an e beam interacting with a Ta foil can generate a high flux of bremsstrahlung radiation.

4.2 Characteristic γ-ray spectrum

The characteristic γ-rays emitted from the Ho sample were measured with the HPGe detector, as mentioned earlier. In our case, the cooling time tc is 30 min. Figure 3A shows the measured characteristic γ-ray spectra for td = 240 min, from which two photopeaks at 73.4 keV (Ig = 1.88%) and 91.4 keV (Ig = 2.30%) are clearly observed, whereas the 37.3 keV γ-ray for 164mHo does not appear. This is because the amount of 164mHo produced during a single-shot irradiation was not enough to be detectable after the cooling time of 30 min. However, the signal and background counts around the γ-ray energy at 37.3 keV can be used to safely determine an upper limit for the yield of 164mHo. By integrating over the energy region of the characteristic γ-ray at 37.3 keV, the signal and background counts are 242, which gives an upper limit of 0.85 × 106 for the 164mHo yield according to Eq. 3a.

FIGURE 3
www.frontiersin.org

FIGURE 3. Typical γ-ray spectrum from the activated Ho foils (tc = 30 min and td = 240 min) (A) and accumulated photopeak counts as a function of detection time (B).

As mentioned previously, reasonably partitioning the Cg is vital to obtain the px and the resulting IR. The Cg values for 73.4 and 91.4 keV lines as functions of td are shown in Figure 3B. The temporal variation of Cg can be fitted well with Eq. 3b. The fitting curve can be re-written as

Cg=P0eλgtd+P1eλmtd+P2,(8)

where P0, P1, and P2 are the coefficients dependent on the px. More specifically, the coefficients P0 and P2 are associated with both the pm and pg. P1 is the function of pm. From Eq. 8, one can see that the Cg value is not only contributed by the ground state of 164Ho but also induced by its isomeric state. In our case, the P1 is fitted to be −1865 ± 229, and the pm value is then obtained to be (0.44 ± 0.05) × 106. The resulting confidence level is approximately 9.0 σ.

4.3 Isomeric ratio calculation

The calculation of the IR relies on the determination of pm and pg. Since the characteristic γ-ray line at 37.3 keV is not observed in our case, we employ the approach shown in Eq. 5 to extract the px values of 164m, gHo. Figure 4A presents 10 groups of px values obtained with the photopeak counts at 91.4 keV. The average pm¯ and pg¯ are calculated to be 0.45 × 106 and 1.48 × 106 per laser shot, respectively. The uncertainty of px is determined by σpx=i=1i=nσi2px/n, where the σipx is the uncertainty of pxi calculated by Eq. 6. Finally, the pm and pg of Ho produced in the experiment are obtained to be (0.45 ± 0.10) × 106 and (1.48 ± 0.14) × 106 per laser shot, respectively. Accordingly, the confidence level of pm is 4.5 σ. It should be noted that the pm value is in good agreement with the one obtained by fitting the Cg curve, as discussed earlier. However, its confidence level is smaller than the confidence level of P1. This is reasonable since the former reasonably considers the error propagation. In addition, the pm value is almost two times lower than the upper limit of 0.85 × 106 mentioned earlier, which in turn validates the feasibility of extracting the 164mHo yield using only the peak counts from the ground-state decay. Similarly, the IR value and its uncertainty σIR are calculated with Eq. 4 and 7, respectively. Figure 4B shows 10 groups of IRs and their average value. The IR value of 164m, gHo is 0.30 ± 0.08, which is less than unity. This is because the spin (J = 6) of the isomeric state is visibly higher than the ground state with J = 1.

FIGURE 4
www.frontiersin.org

FIGURE 4. Ten groups of pm, pg (A), and IRs (B) for the 91.4 keV photopeak. Group No. i (i=1,2,,10) indicates the data calculated with the photopeak counts at tdi =5i+40 min and tdj=5i+120 min. In (A), the blue and red lines stand for the average pm¯ and pg¯, respectively. In (B), the black line represents the average value of IR s.

Photoneutron reaction cross-section of 165Ho calculated with TALYS 1.9 (Koning et al., 2019) and the data from the Experimental Nuclear Reaction Database (EXFOR) are compared and shown in Figure 5A. One can see that the 165Ho(γ, n)164gHo reaction plays a dominant role in the giant dipole resonance (GDR) region. The 165Ho(γ, n)164mHo reaction has a similar distribution with the 165Ho(γ, n)164gHo. However, its maximum cross-section, ∼60 mb, is visibly lower than that of the latter. As the photon energy continues to increase, the multiple emission reactions take over. The maximum cross-section decreases as the number of emitted particles increases. The TALYS calculations are in overall good agreement with the EXFOR data obtained from previous measurements using photon sources caused by positron annihilation in flight (Bergère et al., 1968; Berman et al., 1969). This indicates the reliability of the TALYS inputs and calculations. In addition, it is noticeable that the available experimental data in terms of isomer production are very rare, which can potentially be measured using state-of-the-art laser-Compton scattering facilities generating high-intensity and quasi-monoenergetic γ-ray beams (An et al., 2018; Wang et al., 2022).

FIGURE 5
www.frontiersin.org

FIGURE 5. Calculated (γ, n) cross-sections for 165Ho and the available EXFOR data for comparison (A) and the flux-averaged IR of 164m, gHo in the 165Ho(γ, n) reaction as a function of excitation energy (B). In (A), the experimental data are total cross-sections of the 165Ho(γ, n) and 165Ho(γ, np) reactions, and Snp denotes the threshold energy for the 165Ho(γ, np) reaction. In (B), the previous experimental data (Kolev et al., 1995; Thiep et al., 2011; Do et al., 2013) are obtained by bremsstrahlung photons from RF electron accelerators, and TALYS calculations considering different nuclear level density (NLD) models are also presented for comparison.

The IRs of 164m, gHo can be examined by using different bremsstrahlung radiations from both the laser-plasma accelerator and the RF accelerator. For this purpose, the effective γ-ray energy Eγ from the threshold to the end-point energy can be obtained by using the following relation (Jacobs et al., 1979):

Eγ=EthEmaxφEγσREγEγdEγEthEmaxφEγσREγdEγ,(9)

where the φEγ is calculated with the Geant4 toolkit considering their realistic target arrangements, and σREγ is the cross-section for the 165Ho(γ, n)164m, gHo reaction, which is calculated by using the default option in the TALYS software. In order to understand the effect of excitation energy, the measured IR value of 164m, gHo from the present work, literature values, and TALYS calculations are plotted in Figure 5B as a function of Eγ. From the EXFOR data, one can see that when Eγ > 11 MeV, the IR of 164m, gHo increases with the Eγ and then gets saturated. This is because the input angular momentum brought in by photons is very low. Since the experimental data are not available within the energy range Eγ < 11 MeV, the IRs are further calculated with the TALYS software considering different NLD models. It is found that the IR clearly depends on the Eγ. The increasing and decreasing trends of the IR values appear within the energy range of 8.5 < Eγ <11.0 MeV. Such trends are not only due to the excitation energy effect but also due to the GDR effect. In our experiment, the IR value of 164m, gHo is 0.30 ± 0.08 at Eγ = 12.65 MeV, which is in agreement with both TALYS calculations and the data of Kolev et al. (1995) within the statistical uncertainty. However, the experimental IRs provided by Thiep et al. (2011) and Do et al. (2013) are higher than the TALYS calculations at Eγ > 12 MeV.

5 Conclusion

We carried out the experiment to produce 164m, gHo via photoneutron reaction induced by a laser-accelerated electron beam, in which the IR value of 164m, gHo is determined by using the activation and offline γ-ray spectrometry technique. However, since the characteristic γ-rays from the isomeric decay of 164mHo were not successfully observed, we propose to extract the production yields of 164m, gHo by partitioning counts of photopeak characterizing the ground-state decay. This is different from the approach by extracting the counts of two photopeaks characterizing directly the isomeric and ground states. The production yields of 164m, gHo were successfully extracted to be (0.45 ± 0.10) × 106 and (1.48 ± 0.14) × 106 per laser shot. Accordingly, the IR value is calculated to be 0.30 ± 0.08 at Eγ = 12.65 MeV. The IR as a function of Eγ is further calculated with the TALYS software considering different NLD models. It is found that our result is in agreement with both TALYS calculations and the available experimental data within the statistical uncertainty. In addition, the increasing and decreasing trends of the IR values are observed within the energy range of 8.5 < Eγ <11.0 MeV, suggesting that excitation energy is crucial to determine the IR value of 164m, gHo.

Data availability statement

The original contributions presented in the study are included in the article/Supplementary material; further inquiries can be directed to the corresponding authors.

Author contributions

JZ: Data curation, Formal Analysis, Investigation, Software, Visualization, Writing–original draft. WQ: Data curation, Investigation, Resources, Writing–review and editing. WF: Data curation, Visualization, Writing–review and editing. ZC: Data curation, Software, Writing–review and editing. KL: Data curation, Visualization, Writing–review and editing. CT: Software, Writing–review and editing. XZ: Resources, Writing–review and editing. ZD: Resources, Writing–review and editing. ZZ: Resources, Writing–review and editing. XL: Writing–review and editing. YY: Writing–review and editing. WL: Funding acquisition, Supervision, Writing–review and editing. WZ: Funding acquisition, Supervision, Writing–review and editing.

Funding

The authors declare financial support was received for the research, authorship, and/or publication of this article. This work was supported by the National Key R&D Program of China (Grant No. 2022YFA1603300), the National Natural Science Foundation of China (Grant No. U2230133), the Independent Research Project of the Key Laboratory of Plasma Physics, CAEP (Grant No. JCKYS2021212009), the Open Fund of the Key Laboratory of Nuclear Data, CIAE (Grant No. JCKY2022201C152), the Research Foundation of Education Bureau of Hunan Province, China (No. 22B0453), the Hunan Provincial Natural Science Foundation of China (No. 2023JJ40525), and the Hengyang Municipal Science and Technology Project (No. 202150054076).

Acknowledgments

The authors thank the XingGuang-III operation team for operating the laser system and providing the laser-accelerated electron beam.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Agostinelli, S., Allison, J., Amako, K., Apostolakis, J., Araujoa, H., Arce, P., et al. (2003). GEANT4-a simulation toolkit. Nucl. Instrum. Methods Phys. Res. Sect. A 506 (3), 250–303. doi:10.1016/S0168-9002(03)01368-8

CrossRef Full Text | Google Scholar

An, G., Chi, Y., Dang, Y., Fu, G., Guo, B., Huang, Y., et al. (2018). High energy and high brightness laser Compton backscattering gamma-ray source at IHEP. Matter Radiat. Extrem 3 (4), 219–226. doi:10.1016/j.mre.2018.01.005

CrossRef Full Text | Google Scholar

Bergère, R., Beil, H., and Veyssière, A. (1968). Photoneutron cross sections of La, Tb, Ho and Ta. Nucl. Phys. 121 (2), 463–480. doi:10.1016/0375-9474(68)90433-8

CrossRef Full Text | Google Scholar

Berman, B. L., Kelly, M. A., Bramblett, R. L., Caldwell, J. T., and Fultz, S. C. (1969). Giant resonance in deformed nuclei: photoneutron cross sections for Eu153, Dd160, Ho165, and W186. Phys. Rev. 185 (4), 1576–1590. doi:10.1103/PhysRev.185.1576

CrossRef Full Text | Google Scholar

Bonnet, T., Comet, M., Denis-Petit, D., Gobet, F., Hannachi, M., Tarisien, M., et al. (2013). Response functions of fuji imaging plates to monoenergetic protons in the energy range 0.6–3.2 MeV. Rev. Sci. Instrum. 84 (1), 013508. doi:10.1063/1.4775719

PubMed Abstract | CrossRef Full Text | Google Scholar

Cao, Z., Qi, W., Lan, H., Cui, B., Zhang, X., Deng, Z., et al. (2023). Experimental study of medical isotopes 62,64Cu and 68Ga production using intense picosecond laser pulse. Plasma Phys. control. Fusion 65, 055007. doi:10.1088/1361-6587/acc090

CrossRef Full Text | Google Scholar

Danson, C., Haefner, C., Bromage, J., Butcher, T., Chanteloup, J., Chowdhury, E., et al. (2019). Petawatt and exawatt class lasers worldwide. High. Power Laser Sci. 7, E54. doi:10.1017/hpl.2019.36

CrossRef Full Text | Google Scholar

Do, N. V., Khue, P. D., Thanh, K. T., Hien, N. T., Kim, G., Lee, M., et al. (2013). Measurement of isomeric yield ratios for the natHo(γ, xn)164m, g; 162m, gHo reactions in the bremsstrahlung energy region from 45 to 65 MeV. J. Radioanal. Nucl. 298, 1447–1452. doi:10.1007/s10967-013-2608-6

CrossRef Full Text | Google Scholar

Fan, W., Qi, W., Zhang, J., Cao, Z., Lan, H., Li, X., et al. (2023). Efficient production of nuclear isomer 93mMo with laser-accelerated proton beam and an astrophysical implication on 92Mo production. Available at: https://arxiv.org/ftp/arxiv/papers/2308/2308.02994.pdf.

Google Scholar

Feng, J., Li, Y., Tan, J., Wang, W., Li, Y., Zhang, X., et al. (2022). Laser plasma accelerated ultra-intense electron beam for efficiently exciting nuclear isomers. doi:10.48550/arXiv.2203.06454

CrossRef Full Text | Google Scholar

Günther, M. M., Rosmej, O. N., Tavana, P., Gyrdymov, M., Skobliakov, A., Kantsyrev, A., et al. (2022). Forward-looking insights in laser-generated ultra-intense γ-ray and neutron sources for nuclear application and science. Nat. Commun. 13, 170. doi:10.1038/s41467-021-27694-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Habs, D., and Köster, U. (2011). Production of medical radioisotopes with high specific activity in photonuclear reactions with γ-beams of high intensity and large brilliance. Appl. Phys. B 103, 501–519. doi:10.1007/s00340-010-4278-1

CrossRef Full Text | Google Scholar

Hayakawa, T., Miyamoto, S., Hayashi, Y., Kawase, K., Horikawa, K., Chiba, S., et al. (2008). Half-life of the 164Ho by the (γ,n) reaction from laser Compton scattering γ rays at the electron storage ring NewSUBARU. Phys. Rev. C 74, 065802. doi:10.1103/physrevc.74.065802

CrossRef Full Text | Google Scholar

Hilgers, K., Qaim, S. M., and Sudar, S. (2007). Formation of the isomeric pairs 1³Ndm,g and 11Ndm,g in proton and ³He-particle-induced nuclear reactions. Phys. Rev. C 76, 064601. doi:10.1103/PHYSREVC.76.064601

CrossRef Full Text | Google Scholar

Inagaki, M., Sekimoto, S., Tadokoro, T., Ueno, Y., Kani, Y., Ohtsuki, T., et al. (2020). Production of 99Mo/99mTc by photonuclear reaction using a natMoO3 target. J. Radioanal. Nucl. Chem. 324, 681–686. doi:10.1007/s10967-020-07086-9

CrossRef Full Text | Google Scholar

Jacobs, E., Thierens, H., Frenne, D. D., Clercq, A. D., D'hondt, P., Gelder, P. D., et al. (1979). Product yields for the photofission of 238U with 12-15-20-30-and 70-MeV bremsstrahlung. Phys. Rev. C 19 (2), 422–432. doi:10.1103/PhysRevC.19.422

CrossRef Full Text | Google Scholar

Jonsson, G., and Eriksson, M. (1977). Isomeric ratios in photon-induced spallation reactions at intermediate energies. Z. Phys. A 281, 53–56. doi:10.1007/BF01408612

CrossRef Full Text | Google Scholar

Kim, K., Kim, G., Naik, H., Zaman, M., Yang, S., Song, T., et al. (2015). Excitation function and isomeric ratio of Tc-isotopes from the 93Nb(α, xn) reaction. Nucl. Phys. 935, 65–78. doi:10.1016/j.nuclphysa.2014.12.006

CrossRef Full Text | Google Scholar

Kolev, D., Dobreva, E., Nenov, N., and Todorov, V. (1995). A convenient method for experimental determination of yields and isomeric ratios in photonuclear reactions measured by the activation technique. Nucl. Instrum. Methods Phys. Res. Sect. A 356 (2-3), 390–396. doi:10.1016/0168-9002(94)01319-5

CrossRef Full Text | Google Scholar

Koning, A. J., Rochman, D., Sublet, J., Dzysiuk, N., Fleming, M., and Marck, S. V. D. (2019). TENDL: complete nuclear data library for innovative nuclear science and technology. Nucl. Data Sheets 155, 1–55. doi:10.1016/j.nds.2019.01.002

CrossRef Full Text | Google Scholar

Luo, J., and Jiang, L. (2014). Ground-state and isomeric-state cross sections for 165Ho(n, 2n)164Ho reaction from the reaction threshold to 20 MeV. Phys. Rev. C 89, 014604. doi:10.1103/PhysRevC.89.014604

CrossRef Full Text | Google Scholar

Naik, H., Kim, G. N., Schwengner, R., Kim, K., Zaman, M., Yang, S. C., et al. (2016). Measurement of isomeric ratios for 89g,mZr, 91g,mMo, and 97g,mNb in the bremsstrahlung end-point energies of 16 and 45-70 MeV. Eur. Phys. J. 52, 47. doi:10.1140/epja/i2016-16047-8

CrossRef Full Text | Google Scholar

Nishiuchi, M., Sakaki, H., Dover, N. P., Miyahara, T., Shiokawa, K., Manabe, S., et al. (2020). Ion species discrimination method by linear energy transfer measurement in Fujifilm BAS-SR imaging plate. Rev. Sci. Instrum. 91, 093305. doi:10.1063/5.0016515

PubMed Abstract | CrossRef Full Text | Google Scholar

Peik, E., Schumm, T., Safronova, M. S., Pálffy, A., Weitenberg, J., and Thirolf, P. G. (2021). Nuclear clocks for testing fundamental physics. Quantum Sci. Technol. 6, 034002. doi:10.1088/2058-9565/abe9c2

CrossRef Full Text | Google Scholar

Prelas, M. A., Weaver, C. L., Watermann, M. L., Lukosi, E. D., Schott, R. J., and Wisniewskiet, A. (2014). A review of nuclear batteries. Prog. Nucl. Energ 75, 117–148. doi:10.1016/j.pnucene.2014.04.007

CrossRef Full Text | Google Scholar

Qi, W., Zhang, X., Zhang, B., He, S., Zhang, F., Cui, B., et al. (2019). Enhanced photoneutron production by intense picoseconds laser interacting with gas-solid hybrid targets. Phys. Plasmas. 26, 043103. doi:10.1063/1.5079773

CrossRef Full Text | Google Scholar

Rahman, M. S., Kim, K., Kim, G., Naik, H., Nadeem, M., Hien, N. T., et al. (2016). Measurement of flux-weighted average cross-sections and isomeric yield ratios for 103Rh(γ, xn) reactions in the bremsstrahlung end-point energies of 55 and 60 MeV. Eur. Phys. J. A 52, 194. doi:10.1140/epja/i2016-16194-x

CrossRef Full Text | Google Scholar

Rahman, M. S., Kim, K., Nguyen, T. H., Kim, G., Naik, H., Yang, S., et al. (2020). Measurement of flux-weighted average cross sections of natIn(γ, xn) reactions and isomeric yield ratios of 112mg, 111m,g, 110m,g in with bremsstrahlung. Eur. Phys. J. A 56, 235. doi:10.1140/epja/s10050-020-00245-2

CrossRef Full Text | Google Scholar

Robbie, W., Martin, K., Ross, G., David, C., Rachel, D., Nicholas, B., et al. (2018). Development of focusing plasma mirrors for ultraintense laser-driven particle and radiation sources. Quantum Beam Sci. 2 (1), 1. doi:10.3390/qubs2010001

CrossRef Full Text | Google Scholar

Schlenvoigt, H. P., Haupt, K., Debus, A., Budde, F., Jäckel, O., Pfotenhauer, S., et al. (2008). A compact synchrotron radiation source driven by a laser-plasma wakefield accelerator. Nat. Phys. 4, 130–133. doi:10.1038/nphys811

CrossRef Full Text | Google Scholar

Pan, W., Song, T., Lan, H., Ma, Z., Zhang, J., Zhu, Z., and Luo, W. (2021). Photo-excitation production of medically interesting isomers using intense γ-ray source. Appl Radiat Isotopes 168, 109534. doi:10.1016/j.apradiso.2020.109534

PubMed Abstract | CrossRef Full Text | Google Scholar

Singh, B., and Chen, J. (2018). Nuclear data sheets for A=164. Nucl. Data Sheets 147, 1–381. doi:10.1016/j.nds.2018.01.001

CrossRef Full Text | Google Scholar

Thiep, T. D., An, T. T., Cuong, P. V., Vinh, N. T., and Belov, A. G. (2011). Study of the isomeric ratios in photonuclear reactions of natural holmium and lutetium induced by bremsstrahlungs with endpoint energies in the giant dipole resonance region. J. Radioanal. Nucl. Chem. 290, 515–524. doi:10.1007/s10967-011-1257-x

CrossRef Full Text | Google Scholar

Wang, H., Fan, G., Liu, L., Xu, H., Shen, W., Ma, Y., et al. (2022). Commissioning of laser electron gamma beamline SLEGS at SSRF. Nucl. Sci. Tech. 33, 87. doi:10.1007/s41365-022-01076-0

CrossRef Full Text | Google Scholar

Zilges, A., Balabanski, D., Isaak, J., and Pietralla, N. (2022). Photonuclear reactions-From basic research to applications. Prog. Part. Nucl. Phys. 122, 103903. doi:10.1016/j.ppnp.2021.103903

CrossRef Full Text | Google Scholar

Appendix

The photopeak counts Cx (x=m,g) is the integration of the Activity Ax for x-state considering Ix and εx over a detection time td as follows:

Cx=0tdIxεxAxdt.(A1)

As shown above, the activity Ax is the key factor in the Cx solution. According to Eq. 1 in the proof (on line 204), in irradiation interval, both the productions and the decay properties contribute to the activities of the isomeric and ground states. And in the cooling and the detection interval, only the decay properties of two states contribute to the activities. So, we deduce the activities of the isomeric and ground states for two intervals, i.e., the irradiation interval and the natural decay interval (including cooling interval and detection interval).

Firstly, the number of x-state Nx at the irradiation time tirr can be solved out. In irradiation interval, the Nx changes as a function of tirr and the formula is like to Eq. 1 in the proof. By solving the Eq. 1 in the proof, the Nx at the irradiation time tirr can be obtained as:

Nm=1λmpm1eλmtirr,(A2)
Ng=ηpm1λgλgλmλg1eλmtirrλm1eλgtirr+1λgpg1eλgtirr.(A3)

And then, in the cooling and the detection intervals, the production doesn’t do any contribution to the activities of the isomeric and ground states. Therefore, the Eq. 1 in the proof for the cooling and detection intervals can be written as:

dNmdt=λmNm,dNgdt=ηλmNmλgNg.(A4)

By solving it, the Nx changes as a function of the cooling tc and the detection td times can be obtained as:

Nm=1λmpm1eλmtirreλmtdeλmtm,(A5)
Ng=ηmpmλgλgλmλg1eλmtirreλmtceλmtdλm1eλgtirr×eλgtceλgtd+1λgpg1eλgtirreλgtceλgtd.(A6)

Due to Ax=λxNx, the Ax can be obtained as the following:

Am=pm1eλmtirreλmtdeλmtm,(A7)
Ag=ηmpmλgλmλg1eλmtirreλmtceλmtdλm1eλgtirr×eλgtceλgtd+pg1eλgtirreλgtceλgtd.(A8)

Substituting Eq. 7 into Eq. 1 and solving it, the Cm can be gotten as:

Cm=Imεmpm1eλmtirrλmeλmtc1eλmtd,(A9)

where the Eq. 3a in the proof is deduced out.

And do the same performance to Eq. 8 like Eq. 7. The Cg can be obtained as:

Cg=Igεgpg1eλgtirrλgeλgtc1eλgtd+pmηλgλmλgλm1eλmtirr×eλmtc1eλmtdλmλg1eλgtirreλgtc1eλgtd,(A10)

where the Eq. 3b in the proof is deduced out.

Keywords: isomeric yield ratio, photoneutron reaction, laser-accelerated electron beam, effective γ-ray energy, 164m, gHo

Citation: Zhang J, Qi W, Fan W, Cao Z, Luo K, Tan C, Zhang X, Deng Z, Zhang Z, Li X, Yuan Y, Luo W and Zhou W (2023) Study of the isomeric yield ratio in the photoneutron reaction of natural holmium induced by laser-accelerated electron beams. Front. Astron. Space Sci. 10:1265919. doi: 10.3389/fspas.2023.1265919

Received: 24 July 2023; Accepted: 02 October 2023;
Published: 08 November 2023.

Edited by:

Yi Xu, Horia Hulubei National Institute for Research and Development in Physics and Nuclear Engineering (IFIN-HH), Romania

Reviewed by:

Canel Eke, Akdeniz University, Türkiye
Marina Barbui, Texas A&M University, United States

Copyright © 2023 Zhang, Qi, Fan, Cao, Luo, Tan, Zhang, Deng, Zhang, Li, Yuan, Luo and Zhou. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Wen Luo, wenluo-ok@163.com; Weimin Zhou, zhouwm@caep.cn

These authors share first authorship

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.