logo

Scheme for the excitation of thorium-229 nuclei based on electronic bridge excitation

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Scheme for the excitation of thorium-229 nuclei based on electronic bridge excitation

Lin Li
Zi Li
Chen Wang
Wen-Ting Gan
Xia Hua
Xin Tong
Nuclear Science and TechniquesVol.34, No.2Article number 24Published in print Feb 2023Available online 24 Feb 2023
58500

Thorium-229 possesses the lowest first nuclear excited state, with an energy of approximately 8 eV. The extremely narrow linewidth of the first nuclear excited state, with an uncertainty of 53 THz, prevents direct laser excitation and realization of the nuclear clock. We present a proposal using the Coulomb crystal of a linear chain formed by 229Th3+ ions, where the nuclei of 229Th3+ ions in the ion trap are excited by the electronic bridge (EB) process. The 7P1/2 state of the thorium-229 nuclear ground state is chosen for EB excitation. Using the two-level optical Bloch equation under experimental conditions, we calculate that two out of 36 prepared thorium ions in the Coulomb crystal can be excited to the first nuclear excited state, and it takes approximately 2 h to scan over an uncertainty of 0.22 eV. Taking advantage of the transition enhancement of EB and the long stability of the Coulomb crystal, the energy uncertainty of the first excited state can be limited to the order of 1 GHz.

Coulomb crystalthorium-229electronic bridge transitionisomeric state
1

Introduction

The excitation energy between the nuclear ground state and the first nuclear excited state (isomeric state) of thorium-229 (Th-229) is approximately 8 eV [1-4]. The optical transition linewidth of the first nuclear excited state is 10-4 Hz [5, 6]. The low isomeric energy and narrow linewidth provide the potential for a resonator with a quality of a factor of 1019 and make Th-229 the most suitable choice for the development of nuclear optical clocks [6-8]. Such a Th-229 nuclear clock is expected to be a sensitive probe for the time variation of the fundamental constants of nature and will open opportunities for highly sensitive tests of the fundamental principles of physics, particularly in searches for violations of Einstein’s equivalence principle and new particles [9-11].

The isomeric energy is currently measured in two ways: 1) Measuring the kinetic energy of internal conversion electrons from the Th-229 isomeric state. For example, an electron spectrometer has been used to detect the internal conversion electron energy, and the isomeric energy was obtained as 8.28 ± 0.17 eV [4]. 2) Observing the γ radiation from Th-229 excited nuclei. An isomeric energy of 7.8 ± 0.5 eV was obtained using a magnetic microcalorimeter [2]. Subsequently, a 29.2-keV inter-band excitation of the Th-229 nuclear state was observed using synchrotron radiation [3]. Combined with the in-band 29.2-keV transition observed via nuclear rotational spectroscopy, the Th-229 isomeric state energy was determined to be 8.30 ± 0.92 eV [12]. Recently, a more precise magnetic microcalorimeter with improved energy resolution was developed, and an isomeric energy of 8.10 ± 0.17 eV was obtained [1]. Taking different weights for these measurements, the average value of the isomeric state was determined as 8.12 ± 0.11 eV [13]. Its uncertainty corresponds to 53 THz, which is at least 16 orders of magnitude higher than the narrow linewidth of 10-4 Hz.

To improve the precision of the isomeric energy, synchrotron radiation has been used to irradiate Th-229 doped in a vacuum ultraviolet (VUV) transparent crystal, and γ photons were detected from the decay of the Th-229 excited nuclei. This measurement excluded approximately half of the favored transition search area [14]. Another scheme was proposed to achieve Th-229 nuclei excitation via electron capture and detect the isomeric state of ions in heavy-ion storage rings [15]. Recently, an optical frequency comb was proposed to irradiate a Th-229 dioxide film, and the isomeric energy was measured based on the internal conversion process [16]. Furthermore, a new approach has been suggested to excite the isomeric Th-229 nuclear state via laser-driven electron recollision. The advantage of this approach is that it does not require precise knowledge of the isomeric energy [17].

In this paper, we propose the excitation of the nuclei of 229Th3+ ions in the ion trap via electronic bridge (EB) excitation and compare it with direct nuclear excitation via a pulsed laser. As shown in the inset of Fig. 1, EB excitation starts from an electric dipole (E1) transition, which is utilized to excite the electrons from an initial state to a virtual intermediate state. Subsequently, the intermediate state decays via the magnetic dipole (M1) transition, and the nuclear ground state 5/2+[633] is excited to the isomeric state 3/2+[631] simultaneously. To date, EB schemes for nuclear excitation have been theoretically investigated for 229gTh+ [18], 229Th2+ [19], 229Th3+ [20-22], and 229Th35+ [23]. Our proposed experiment is based on EB excitation in triply charged 229Th3+ ions. Owing to their simple electronic energy levels, the trapped 229Th3+ ions can be laser-cooled to form a linear-chain Coulomb crystal using two closed transitions in the nuclear ground state of 229Th3+ (229gTh3+) [24]. When the 229gTh3+ ions are prepared in the 7P1/2 state, EB excitation is triggered by a 350-nm laser, and isomeric states are expected to be populated. Finally, the isomeric state is detected based on the isomer shift of the cooling lasers [7].

Fig. 1
Diagram of electronic energy levels and electric-dipole transitions of Th-229 triply charged ions. The insert shows the processes of EB excitation, where the red arrows are the 150-nm M1 transition. The black horizontal lines indicate the nuclear ground state of Th-229, and the red horizontal lines indicate the isomeric state of Th-229. The purple arrow is the 350-nm EB excitation. The lifetime of the electronic excited state is indicated in the parentheses. The optical transition wavelengths are in nm, and the integers near the atomic levels indicate the principal quantum numbers.
pic
2

Preparation of 229gTh3+ ion Coulomb crystals and the 7P1/2 state

229gTh3+ ions are typically produced by laser ablation [24, 25]. The linear Paul trap used to trap 229gTh3+ ions are described in detail in our previous papers [26, 27]. The trapped 229gTh3+ ions are laser-cooled to form a Coulomb crystal.

The energy levels of laser cooling and fluorescence detection of the trapped ions are shown in Fig. 2. The 229gTh3+ ions are Doppler-laser-cooled on a 52F5/2 62D3/2 transition at 1088 nm. Alternatively, a closed optical transition of the three-level Λ system, 52F5/2 62D5/2 52F7/2, can be used to laser-cool the 229gTh3+ ions with two lasers at 690 nm and 984 nm [7]. The population in the three hyperfine levels of |5F5/2,F=3,4,5 can be transferred to |5F5/2,F=1 and |5F5/2,F=2 via the 62D3/2 state, and the population in the |5F7/2,F=1 state is cumulated as shown in Fig. 2. Therefore, the ions at other hyperfine levels are transferred to the closed system of 52F5/2 62D5/2 52F7/2, and the second laser cooling cycle is realized.

Fig. 2
State preparation scheme of 229gTh3+ ions. The cooling scheme utilizes the two closed system indicated by the black solid arrows. The wavy arrows indicate spontaneous radiation, and the blue solid arrows indicate the transition to the 7P1/2 state. The red dashed arrows are the transitions to transfer ions at |5F5/2,F=3,4,5 to |5F5/2,F=1 and |5F5/2,F=2. The black dashed arrows are the transitions to transfer ions at |5F7/2,F=2,3,4,5,6 to |5F7/2,F=1. The integers near the atomic levels indicate the total angular momentum quantum numbers F.
pic

Once the 229gTh3+ ions are laser-cooled to 50 μK, the Doppler broadening of the transition 62D5/2 52F7/2 at 984 nm is approximately 1 MHz [24]. At this temperature, the corresponding Doppler broadenings of direct nuclear excitation at 150 nm and EB excitation at 350 nm are estimated to be 6.5 MHz and 2.8 MHz, respectively. The shapes of Coulomb crystals can be manipulated into a linear chain by adjusting the radiofrequency (RF) and endcap voltages [28]. Two-dimensional and linear-chain Coulomb crystals can be simulated with single-ion resolution [29, 30]. Using the GPU-accelerated LAMMPS [31] wrapped by the LIon package [32], a Coulomb crystal consisting of 36 229gTh3+ ions are simulated, as shown in Fig. 3. When an endcap voltage of 0.2 V and an RF voltage (200 V0-pk) of 2 MHz are applied to the ion trap, a linear-chain Coulomb crystal of 36 229gTh3+ ions (Fig. 3(c)) is obtained. When an RF voltage (200 V0-pk) of 2 MHz and endcap voltage of 0.025 V are applied to the ion trap, a linear-chain Coulomb crystal of 100 229gTh3+ ions is obtained. From the simulation results, the motion amplitude of ions in the x- and y-directions (radial direction) is known to be less than 0.6 μM when the temperature is below 5 mK, which is consistent with the motion amplitude of single ions [33]. Therefore, a strongly focused laser spot size of over 1.2 μM is required to increase the EB and direct nuclear excitation rates.

Fig. 3
Simulation results of a Coulomb crystal consisting of 36 229gTh3+ ions. With the same RF voltages (200 V0-pk), the Coulomb crystals’ shapes are changed under different endcap voltages (a) 0.4 V, (b) 0.3 V, and (c) 0.2 V. See text for details.
pic

To induce EB excitation from the 7P1/2 state of the 229gTh3+ ions, the three-level transition 52F5/2 62D3/2 72P1/2 can be used to accumulate the population at the 7P1/2 state. Once the 229gTh3+ ions are laser-cooled to 50 μK, the 690-nm and 984-nm lasers are turned off in sequence to maintain the ions at the 52F5/2 state, and the 1088-nm laser is kept on. To transfer a large population to the 72P1/2 state with a lifetime of 1 ns, both the 269-nm and 196-nm transitions are excited simultaneously, as shown in Fig. 4. The natural linewidths of the 269-nm and 196-nm transitions are approximately 2π ⋅ 32 MHz and 2π ⋅ 127 MHz, respectively [34]. Continuous wave (CW) lasers at both wavelengths are commercially available [35].

Fig. 4
Expected number of excited nuclei is calculated as a function of time when 100 triply charged Th-229 ions are directly illuminated by a direct excitation laser.
pic

The spontaneous radiation rate of the hyperfine structure level transitions between two electronic energy levels is [36] Γfi=|ξfJfIgFfdξiJiIgFi|24ω33c3, (1) where ξi (ξf) incorporates all other electronic quantum numbers, Ji (Jf) is the total electronic quantum number of the initial (final) state, Fi (Ff) is the total quantum number of the initial (final) state, Ig is the nuclear spin of the nuclear ground state, d is the electron dipole, ω is the angular frequency of the electronic transition, is the reduced Planck constant, and c is the speed of light in a vacuum. |JfIgFfdJiIgFi|2 is the transition strength of the hyperfine structure level, which can be expressed as |ξfJfIgFfdξiJiIgFi|2=(2Fi+1)(2Ff+1){JfFfIgFiJi1}2|JfdJi|2, (2) where |JfdJi|2 denotes the transition strength of the electronic energy levels. The spontaneous radiation rate of the electronic energy levels is Γspont=|ξfJfdξiJi|24ω33c3. (3) The excitation rate of the hyperfine structure levels between the two electronic energy levels is [37] Γexc=Γfiπ2c2ω3Pω2Ff+12Fi+1, (4) where denotes the laser spectral intensity. The stimulated emission rate of the hyperfine structure levels between the two electronic energy levels is Γsti=Γexc2Fi+12Ff+1. (5) With a laser power of 1 mW and 100-μM laser spot sizes for 1088 nm, 269 nm, and 196 nm lasers, the steady-state condition can be fulfilled for all 52F5/2, 72S1/2, 62D3/2, and 72P1/2 states under continuous illumination. For multiple hyperfine structure transitions between electronic transitions 52F5/2 62D3/2 and 72S1/2 72P1/2, the power of the lasers is divided equally. Thus, after state preparation, the population of the |7P1/2,F=2 state accounts for approximately 12% of the total number of trapped ions.

3

Calculation of the population in the nuclear excited state

The time-dependent nuclear excitation probability ρexc(t) for a single nucleus under resonant irradiation is given by Torrey’s solution to the optical Bloch equations as follows [38-40]: Ωeg>|ΓΓ˜|2ρexc(t)=Ωeg22(ΓΓ˜+Ωeg2)[1e12(Γ+Γ˜)t×(cos(λt)+Γ+Γ˜2λsin(λt))], (6) Ωeg<|ΓΓ˜|2ρexc(t)=Ωeg22(ΓΓ˜+Ωeg2)[1e12(Γ+Γ˜)t×(cosh(λt)+Γ+Γ˜2λsinh(λt))]. (7) Here, Ωeg denotes the Rabi frequency between the isomeric and nuclear ground states. Γ=Γγnr is the total linewidth of the isomeric state, and Γγ denotes the γ decay rate of the isomeric state. Because the ionization energy of 229gTh3+ (28.6 eV) is considerably greater than the energy of the isomeric state, the internal conversion process is strongly forbidden [41]. The possible non-radiative transition linewidth Γnr is dominated by the EB transition for 229gTh3+. Γ˜=Γ+ΓL2+Γadd denotes the total transition linewidth [40], ΓL denotes the laser linewidth used for nuclear excitation, and Γadd is the additional incoherent linewidth, such as phonon coupling of ions in a Coulomb crystal. Because Γadd is significantly smaller than the laser linewidth ΓL, the influence of the incoherent linewidth is negligible. t denotes the interaction time between the laser and 229gTh3+ ions. Here, λ=|Ωeg2(ΓΓ˜)2/4|.

Under the resonant condition, the Rabi frequency between the isomeric state and nuclear ground state is [39] Ωeg=2πc2ICeg2Γωm3. (8) Here, I is the intensity of the excitation laser, and ωm is the angular frequency of the nuclear transition. The Zeeman splitting caused by the geomagnetic field between the two outermost lines is below 1 kHz and is negligible. Therefore, the Clebsch–Gordan coefficient of the sub-state transition between the isomeric state and the nuclear ground state Ceg is 1 [16].

Assuming that the intensities of all lasers have a Gaussian distribution, the Rabi frequency can be modified via the following two situations:

(i) If the laser linewidth (ΓL) is greater than the Doppler-broadening linewidth (ΓD), only part of the laser with a matching frequency range can effectively interact with the triply charged thorium ions. The equivalent Doppler-broadening linewidth is ΓD'=12πln2ΓD. (9) The effective laser intensity is Ieff=IΓD'2ΓD'2gL(ω)dω. (10) The modified Rabi frequency is Ωeg'=2πc2IeffCeg2Γωm3. (11)

(ii) If the laser linewidth (ΓL) is smaller than the Doppler-broadening linewidth (ΓD), only part of the ions interact with the laser light at the frequency resonance at a certain time, and the equivalent laser linewidth is ΓL'=12πln2ΓL. (12) The effective Doppler-broadening linewidth is Γeff=ΓΓL'2ΓL'2gD(ω)dω. (13) The modified Rabi frequency is Ωeg'=2πc2ICeg2Γeffωm3. (14)

4

Direct nuclear excitation

The Th-229 isomeric energy is 8.12 ± 0.11 eV, corresponding to a wavelength of 150.7-154.8 nm. Currently, available light sources such as the VUV pulsed laser [42], synchrotron radiation light source [14], and the 7th harmonic of a Yb-doped fiber laser [16, 43-45] are suitable candidates for direct nuclear excitation. Here, based on resonance-enhanced four-wave mixing, the parameters of a tunable and pulsed VUV laser source are a pulse energy of EL = 13.2  μJ around 150 nm, a bandwidth of δνL = 15 GHz, and a repetition rate of RL = 10 Hz, which are used to calculate the direct nuclear excitation rate [42].

Assuming a strongly focused laser with a 3-μm spot radius to irradiate the linear chain of 100 ions, the corresponding laser intensity I=ELRLπr2=4.67×106 W/m2 can be obtained. The temperature of Th3+ ions is 50 μK, and the corresponding Doppler-broadening linewidth ΓD = 6.5 MHz is narrower than the laser linewidth. Therefore, the modified Rabi frequency at resonance Ωeg'(δ=0) = 107 Hz can be obtained based on Eq. (9)-(11), which is significantly smaller than the total transition linewidth Γ˜ shown in Table 1. In this case, Ωeg'<|ΓΓ˜|2 is satisfied. The number of excited nuclei is calculated as a function of time using Eq. (7). As shown in Fig. 1, the number of excited Th-229 nuclei is only 0.12 out of 100 Th-229 ions and reaches saturation at an irradiation time of 30000 s. The required laser irradiation time is beyond the stable trapping time; therefore, no Th-229 isomeric state can be detected under the current experimental conditions.

Table 1
Values of variables used for the calculation of the number of excited nuclei using Eq. (7).
Variable Description Value Comment
Γ˜ Total transition linewidth 2 π⋅ 7.5 GHz Significantly larger than the nuclear transition linewidth
Γγ Radiative linewidth 10-4 Hz Estimated from theory
ΓD Doppler-broadening linewidth 6.5 MHz Doppler broadening of ions from direct nuclear excitation
N Number of ions 100 Laser-cooled ions
Ωeg Rabi frequency 1156 Hz Direct excitation between the isomeric and nuclear ground states
Ωeg'(δ=0) Modified Rabi frequency 107 Hz Modified Rabi frequency at resonance
Show more
5

Electronic bridge excitation

Direct nuclear excitation is limited by the natural linewidth of the isomeric state and the power of the excitation laser. If EB excitation occurs, the transition linewidth can be increased by approximately 40 times compared with direct nuclear excitation [46]. EB excitation can be utilized by exciting the quantum state of the electronic state 7P1/2 of 229gTh3+ (7P1/2 5/2+[633]) to the electronic state 7S1/2 of 229mTh3+ (7S1/2 3/2+[631]), as shown in Fig. 1. The prepared 229gTh3+ Coulomb crystal comprising a linear chain of 36 ions is shown in Fig. 3(c), among which there are four ions (12% of the total number of trapped ions) in the 7P1/2 state under steady-state conditions (see Sect. 2).

The wavelength of the laser used for EB excitation is 350 nm. A 350-nm laser with a power of 250 mW can be obtained from Toptica DLC TA-SHG pro and focused to a spot size of 10  μmeter. Thus, an intensity of 7.95 × 108 Wm-2 can be achieved. Therefore, the Rabi frequency is estimated as 334 kHz.

The EB laser linewidth, typically 500 kHz at 350 nm, is narrower than the Doppler broadening of the trapped thorium ions. The modified Rabi frequency Ωeg' is 140 kHz at resonance and 98 kHz upon the detuning ΓD/2 (shown in Table 2). Then, the condition Ωeg'<|ΓΓ˜|2 is satisfied, and thus the number of excited nuclei as a function of time can be obtained based on Eq. (7), as shown in Fig. 5. If the excitation energy from the 7P1/2 state of 229gTh3+ to the 7S1/2 state of 229mTh3+ is at resonance, approximately 2 229gTh3+ ions can be excited to the isomeric state after 300-μs irradiation by the 350-nm laser. If the excitation laser is detuned by ΓD/2, approximately 1.8 229gTh3+ ions can be excited to the isomeric state within the same irradiation time.

Table 2
Values of variables used for the calculation of the number of excited nuclei using Eq. (7).
Variable Value Comment
ΓL 2π⋅ 500 kHz Larger than the nuclear transition linewidth
Γγ 10-4 Hz Estimated from theory
ΓD 2.8 MHz Doppler broadening of ions from EB excitation
N 36 Laser-cooled ions
αeb 40 EB decay of Th3+ ions
Ωeg 334 kHz Rabi frequency
Ωeg'(δ=0) 140 kHz Modified Ωeg at resonance
Ωeg'(δ=ΓD2) 98 kHz Modified Ωeg detune ΓD2
Show more
Fig. 5
Expected number of excited nuclei is calculated as a function of time when 36 triply charged Th-229 ions are trapped and four ions are at the 7P1/2 state.
pic

After successful EB excitation, Th-229 ions in the isomeric state can be identified based on the different hyperfine structures of the isomeric and nuclear ground states. Detailed hyperfine structures of the isomeric state are shown in Appendix B. After EB nuclear excitation, the EB laser at 350 nm and the state-preparation lasers at 196 nm and 269 nm are all turned off, and the cooling lasers at 690 nm and 984 nm are immediately switched back on and combined with the 1088-nm laser to cool the ions. The nuclear-excited ions remain in the isomeric states and accumulate in the electronic state of 52F5/2, as illustrated in Fig. 6. The remaining nuclear-unexcited ions remain in the nuclear ground state of the Doppler cooling cycle of 52F5/2, 62D5/2, 52F7/2 and 62D3/2. The isomer shifts of the 6D5/2 and 6D3/2 electronic states are both approximately 400 MHz [24, 47]. The 690-nm and 984-nm lasers are off-resonance for the Doppler cooling of ions in the isomeric state. As a result, the nuclear-excited 229mTh3+ ions leave the 690-nm and 984-nm lasers’ cooling cycle and appear as dark ions in the Coulomb crystal [7]. In the linear-chain 229Th3+ ion Coulomb crystal, single ions are distinguishable, and any dark 229mTh3+ ions generated by a successful EB excitation event can be detected.

Fig. 6
Experimental steps and the time of each laser switch.
pic

The current reported uncertainty on the isomeric state energy is 0.22 eV [13]. To cover this uncertainty range, approximately 1.9 × 107 scans are required for each scanning interval of the Doppler-broadening linewidth ΓD. The time required for a single scan step is approximately 300 μs; therefore, the total scan time is approximately 5670 s, as shown in Table 3. The final isomeric energy is determined by the sum of the energy intervals between the electronic state 7P1/2 of 229gTh3+ and the electronic state 7S1/2 of 229mTh3+ and the photon energy of the 350-nm laser. Because the uncertainty on the natural linewidth of the 7P1/2 state of 229gTh3+ is 1 GHz, which is significantly larger than the laser and Doppler-broadening linewidths, the uncertainty on the isomeric energy is of the order of 1 GHz (corresponding to 4 × 10-6 eV), dominated by the natural linewidth of the 7P1/2 state.

Table 3
Main parameters for 350-nm irradiation on 36 trapped ions.
Variable Value
Time per scan step 300 μs
Number of ions at the 7P1/2 state 4
Number of laser-cooled ions 36
Number of isomeric states per scan step 2(δ=0)
Number of isomeric states per scan step 1.8(δ=ΓD2)
Time required to scan 0.22 eV 5670 s
Show more
6

Conclusion

In this paper, considering the Doppler broadening of thorium ions in the ion trap, we show that 2 229gTh3+ ions out of 36 trapped ions can be excited to the isomeric state under resonant conditions via EB excitation. We propose that a total measurement time of approximately 2 h can be achieved for the current uncertainty on the isomeric state energy of 0.22 eV. If a cryogenic linear Paul trap is used, the ions can be stably trapped for a longer period and the radiative lifetime of the isomeric state can be measured. The utilization of EB excitation can reduce the uncertainty on the Th-229 isomeric energy to approximately 1 GHz. The calculation is based on the theoretical prediction that the transition linewidth of EB excitation is 40 times larger than that of direct nuclear excitation. If EB excitation is not observed, as proposed in this paper, it may indicate that the coupling between the electrons and the nuclear core is not as strong as expected or that the explored nuclear transition energy range is not accurate, and the laser energy used for the EB excitation needs to be increased. However, the successful observation of EB excitation will further improve the accuracy of isomeric energy, pave the way for the development of a nuclear optical clock, test the temporal variation of fundamental constants, and provide new methods for studying nuclear physics.

References
[1] S. Tomas, G. Jeschua, H. Daniel et al.,

Measurement of the Th-229 isomer energy with a magnetic microcalorimeter

. Phys. Rev. Lett. 125, 142503 (2020). doi: 10.1103/PhysRevLett.125.142503
Baidu ScholarGoogle Scholar
[2] B. R. Beck, J. A. Becker, P. Beiersdorfer et al.,

Energy splitting of the ground-state doublet in the nucleus 229Th

. Phys. Rev. Lett. 98, 142501 (2007). doi: 10.1103/PhysRevLett.98.142501
Baidu ScholarGoogle Scholar
[3] A. Yamaguchi, H. Muramatsu, T. Hayas et al.,

Energy of the 229Th nuclear clock isomer determined by absolute gamma-ray energy difference

. Phys. Rev. Lett. 123, 222501 (2019). doi: 10.1038/s41586-019-1533-4
Baidu ScholarGoogle Scholar
[4] B. Seiferle, L. von der Wense, P. V. Bilous et al.,

Energy of the 229Th nuclear clock transition

. Nature 573, 243-246 (2019). doi: 10.1038/s41586-019-1533-4
Baidu ScholarGoogle Scholar
[5] N. Minkov and A. Palffy,

Theoretical predictions for the magnetic dipole moment of 229mTh

. Phys. Rev. Lett. 122, 162502 (2019). doi: 10.1103/PhysRevLett.122.162502
Baidu ScholarGoogle Scholar
[6] L. von der Wense, B. Seiferle,

The 229Th isomer: prospects for a nuclear optical clock

. Eur. Phys. J. A 56, 277 (2020). doi: 10.1140/epja/s10050-020-00263-0
Baidu ScholarGoogle Scholar
[7] E. Peik and Chr.Tamm,

Nuclear laser spectroscopy of the 3.5 eV transition in Th-229

. EPL 61, 181 (2003). doi: 10.1209/epl/i2003-00210-x
Baidu ScholarGoogle Scholar
[8] C. J. Campbell, A. G. Radnaev, A. Kuzmich et al.,

Single-ion nuclear clock for metrology at the 19th decimal place

. Phys. Rev. Lett. 108, 120802 (2012). doi: 10.1103/PhysRevLett.108.120802
Baidu ScholarGoogle Scholar
[9] V.V. Flambaum,

Enhanced effect of temporal variation of the fine structure constant and the strong interaction in 229Th

. Phys. Rev. Lett. 97, 092502 (2006). doi: 10.1103/PhysRevLett.97.092502
Baidu ScholarGoogle Scholar
[10] E. Peik, T. Schumm, M. Safronova et al.,

Nuclear clocks for testing fundamental physics

. Quantum Sci. Technol. 6, 034002 (2021). doi: 10.1088/2058-9565/abe9c2
Baidu ScholarGoogle Scholar
[11] P. Fadeev, J. Berengut, V. Flambaum,

Sensitivity of Th-229 nuclear clock transition to variation of the fine-structure constant

. Phys. Rev. A 102, 052833 (2020). doi: 10.1103/PhysRevA.102.052833
Baidu ScholarGoogle Scholar
[12] T. Masuda, A. Yoshimi, A. Fujieda et al.,

X-ray pumping of the 229Th nuclear clock isomer

. Nature 573, 238242 (2019). doi: 10.1038/s41586-019-1542-3
Baidu ScholarGoogle Scholar
[13] L. von der Wense,

Ticking toward a nuclear clock

. Physics 13, 152 (2020). doi: 10.1103/Physics.13.152
Baidu ScholarGoogle Scholar
[14] J. Jeet, C. Schneider, S. T. Sullivan et al.,

Results of a direct search using synchrotron radiation for the low-energy 229Th nuclear isomeric transition

. Phys. Rev. Lett. 114, 253001 (2015). doi: 10.1103/PhysRevLett.114.253001
Baidu ScholarGoogle Scholar
[15] X. Ma, W. Wen, Z. Huang et al.,

Proposal for precision determination of 7.8 eV isomeric state in 229Th at heavy ion storage ring

. Phys. Scr. 2015, 014012 (2015). doi: 10.1088/0031-8949/2015/T166/014012
Baidu ScholarGoogle Scholar
[16] L. von der Wense, C.K. Zhang,

Concepts for direct frequency-comb spectroscopy of 229mTh and an internal-conversion-based solid-state nuclear clock

. Eur. Phys. J. D 74, 146 (2020). doi: 10.1140/epjd/e2020-100582-5
Baidu ScholarGoogle Scholar
[17] W. Wang, J. Zhou, B.Q. Liu et al.,

Exciting the Isomeric 229Th Nuclear State via Laser-Driven Electron Recollision

. Phys. Rev. Lett. 127, 052501 (2021). doi: 10.1103/PhysRevLett.127.052501
Baidu ScholarGoogle Scholar
[18] S.G. Porsev, V.V. Flambaum,

Electronic bridge process in 229Th+

. Phys. Rev. A 81, 042516 (2010). doi: 10.1103/PhysRevA.81.042516
Baidu ScholarGoogle Scholar
[19] E. Peik, M. Okhapkin,

Nuclear clocks based on resonant excitation of γ-transitions

. CR. Phys. 16, 516-523 (2015). doi: 10.1016/j.crhy.2015.02.007
Baidu ScholarGoogle Scholar
[20] S.G. Porsev, V.V. Flambaum,

Effect of atomic electrons on the 7.6-eV nuclear transition in 229Th3+

. Phys. Rev. A 81, 032504 (2010). doi: 10.1103/PhysRevA.81.032504
Baidu ScholarGoogle Scholar
[21] P.V. Bilous, E. Peik, A. Pálffy,

Laser-induced electronic bridge for characterization of the 229mTh → 229gTh nuclear transition with a tunable optical laser

. New J Phys. 20, 013016 (2018). doi: 10.1088/1367-2630/aa9cd9
Baidu ScholarGoogle Scholar
[22] N.Q. Cai, G.Q. Zhang, C.B. Fu et al.,

Populating 229mTh via two-photon electronic bridge mechanism

. Nucl. Sci. Tech. 32, 59 (2021). doi: 10.1007/s41365-021-00900-3
Baidu ScholarGoogle Scholar
[23] P.V. Bilous, H. Bekker, J.C. Berengut et al.,

Electronic bridge excitation in highly charged 229Th ions

. Phys. Rev. Lett. 124, 192502 (2020). doi: 10.1103/PhysRevLett.124.192502
Baidu ScholarGoogle Scholar
[24] C.J. Campbell, A.G. Radnaev, A. Kuzmich,

Wigner crystals of 229Th for optical excitation of the nuclear isomer

. Phys. Rev. Lett. 106, 223001 (2011). doi: 10.1103/PhysRevLett.106.223001
Baidu ScholarGoogle Scholar
[25] C.J. Campbell, A.V. Steele, L.R. Churchill et al.,

Multiply charged thorium crystals for nuclear laser spectroscopy

. Phys. Rev. Lett. 102, 233004 (2009). doi: 10.1103/PhysRevLett.102.233004
Baidu ScholarGoogle Scholar
[26] H.X. Li, Y. Zhang, S.G. He et al.,

Determination of the geometric parameters κz and κr of a linear Paul trap

. Chinese J Phys. 60, 61-67 (2019). doi: 10.1016/j.cjph.2019.03.017
Baidu ScholarGoogle Scholar
[27] M. Li, Y. Zhang, Q.Y. Zhang et al.,

An efficient method for producing 9Be+ ions using a 2+ 1 resonance-enhanced multiphoton ionization process

. J. Phys. B At. Mol. Opt. Phys. 55, 035002 (2022). doi: 10.1088/1361-6455/ac4c8f
Baidu ScholarGoogle Scholar
[28] H. C. Nägerl, W. Bechter, J. Eschner et al.,

Ion strings for quantum gates

. Appl. Phys. B 66, 603608 (1998). doi: 10.1007/s003400050443
Baidu ScholarGoogle Scholar
[29] K. Okada, M. Wada, T. Takayanagi et al.,

Characterization of ion Coulomb crystals in a linear Paul trap

. Phys. Rev. A 81, 013420 (2010). doi: 10.1103/PhysRevA.81.013420
Baidu ScholarGoogle Scholar
[30] M.T. Bell, A.D. Gingell, J.M. Oldham et al.,

Ion-molecule chemistry at very low temperatures: cold chemical reactions between Coulomb-crystallized ions and velocity-selected neutral molecules

. Faraday Discuss. 142, 73-91 (2009). doi: 10.1039/b818733a
Baidu ScholarGoogle Scholar
[31] S. Plimpton,

Fast parallel algorithms for short-range molecular dynamics

. J. Comput. Phys. 117, 119 (1995). doi: 10.1006/jcph.1995.1039
Baidu ScholarGoogle Scholar
[32] E. Bentine, C. J. Foot, D. Trypogeorgos,

(py)LIon: A package for simulating trapped ion trajectories

. Comput. Phys. Commun. 253, 107187 (2020). doi: 10.1016/j.cpc.2020.107187
Baidu ScholarGoogle Scholar
[33] P. Richerme,

Two-dimensional ion crystals in radio-frequency traps for quantum simulation

. Phys. Rev. A 94, 032320 (2016). doi: 10.1103/PhysRevA.94.032320
Baidu ScholarGoogle Scholar
[34] U.I. Safronova, W.R. Johnson, M.S. Safronova,

Excitation energies, polarizabilities, multipole transition rates, and lifetimes of ions along the francium isoelectronic sequence

. Phys. Rev. A 76, 042504 (2007). doi: 10.1103/PhysRevA.76.042504
Baidu ScholarGoogle Scholar
[35] L. Kang, Z. Lin,

Deep-ultraviolet nonlinear optical crystals: concept development and materials discovery

. Sci. Appl. 11, 201 (2022). doi: 10.1038/s41377-022-00899-1
Baidu ScholarGoogle Scholar
[36] M. Auzinsh, D. Budker, S. Rochester, Optically polarized atoms: understanding light-atom interactions. 1nd edn. (Oxford, New York, 2010), pp. 141144
[37] I. I. Sobelman, Atomic spectra and radiative transitions. 2nd edn. (Springer Berlin, Heidelberg, 1992), pp. 203204
[38] H. C. Torrey,

Transient nutations in nuclear magnetic resonance

. Phys. Rev. 76, 1059 (1949). doi: 10.1103/PhysRev.76.1059
Baidu ScholarGoogle Scholar
[39] P. J. Colmenares, J. L. Paz,

On the analytical solution of the optical Bloch equations

. Opt. Commun. 284, 51715176 (2011). doi: 10.1016/j.optcom.2011.07.008
Baidu ScholarGoogle Scholar
[40] L. von der Wense, B. Seiferle, P. V. Bilous et al.,

The theory of direct laser excitation of nuclear transitions

. Eur. Phys. J. A 56: 176 (2020). doi: 10.1140/epja/s10050-020-00177-x
Baidu ScholarGoogle Scholar
[41] P.F.A. Klinkenberg,

Spectral structure of trebly ionized thorium, Th IV

. Physica. B+C 151, 552-567 (1988). doi: 10.1016/0378-4363(88)90312-9
Baidu ScholarGoogle Scholar
[42] S. J. Hanna, P. Campuzano-Jost, E. A. Simpson et al.,

A new broadly tunable (7.4-10.2eV) laser based VUV light source and its first application to aerosol mass spectrometry

. Int. J. Mass Spectrom. 279, 134-146 (2009). doi: 10.1016/j.ijms.2008.10.024
Baidu ScholarGoogle Scholar
[43] C. Zhang, S.B. Schoun, C.M. Heyl et al.,

Noncollinear enhancement cavity for record-high out-coupling efficiency of an extreme-UV frequency comb

. Phys. Rev. Lett. 125, 093902 (2020). doi: 10.1103/PhysRevLett.125.093902
Baidu ScholarGoogle Scholar
[44] A. Cingoz, D.C. Yost, T.K. Allison et al.,

Direct frequency comb spectroscopy in the extreme ultraviolet

. Nature 482, 6871 (2012). doi: 10.1038/nature10711
Baidu ScholarGoogle Scholar
[45] I. Pupeza, S. Holzberger, T. Eidam et al.,

Compact high-repetition-rate source of coherent 100 eV radiation

. Nat. Photon. 7, 608612 (2013). doi: 10.1038/nphoton.2013.156
Baidu ScholarGoogle Scholar
[46] R. A. Müller, A. V. Volotka, S. Fritzsche et al.,

Theoretical analysis of the electron bridge process in 229Th3+

. Nucl. Instrum. Meth. B 408, 8488 (2017). doi: 10.1016/j.nimb.2017.05.004
Baidu ScholarGoogle Scholar
[47] J. Thielking, M. V. Okhapkin, P. Głowacki et al.,

Laser spectroscopic characterization of the nuclear-clock isomer 229mTh

. Nature 556, 321325 (2018). doi: 10.1038/s41586-018-0011-8
Baidu ScholarGoogle Scholar
[48] M. S. Safronova, U. I. Safronova, A. G. RadnaeM et al.,

Magnetic dipole and electric quadrupole moments of the 229Th nucleus

. Phys. Rev. A 88, 060501(R) (2013). doi: 10.1103/PhysRevA.88.060501
Baidu ScholarGoogle Scholar
[49] J.C. Berengut, V. A. Dzuba, V.V. Flambaum et al.,

Proposed Experimental Method to Determine α Sensitivity of Splitting between Ground and 7.6 eV Isomeric States in 229Th

. Phys. Rev. Lett. 102, 210801 (2009). doi: 10.1103/PhysRevLett.102.210801
Baidu ScholarGoogle Scholar