This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Brought to you by:
Paper The following article is Open access

Performance of a 229Thorium solid-state nuclear clock

, , , , , , and

Published 17 August 2012 © IOP Publishing and Deutsche Physikalische Gesellschaft
, , Citation G A Kazakov et al 2012 New J. Phys. 14 083019 DOI 10.1088/1367-2630/14/8/083019

1367-2630/14/8/083019

Abstract

The 7.8 eV nuclear isomer transition in 229thorium has been suggested as a clock transition in a new type of optical frequency standard. Here we discuss the construction of a 'solid-state nuclear clock' from thorium nuclei implanted into single crystals transparent in the vacuum ultraviolet range. We investigate crystal-induced line shifts and broadening effects for the specific system of calcium fluoride. At liquid nitrogen temperatures, the clock performance will be limited by decoherence due to magnetic coupling of the thorium nuclei to neighboring nuclear moments, ruling out the commonly used Rabi or Ramsey interrogation schemes. We propose clock stabilization based on a fluorescence spectroscopy method and present optimized operation parameters. Taking advantage of the large number of quantum oscillators under continuous interrogation, a fractional instability level of 10−19 might be reached within the solid-state approach.

Export citation and abstract BibTeX RIS

Content from this work may be used under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Atomic clocks have a significant impact both on fundamental science and on everyday life technology. Prominent examples of the latter are satellite-based navigation systems, such as GPS, GLONASS and GALILEO, and telecommunication networks. On the fundamental side, the current definition of the unit of time (SI second) is a duration corresponding to 9192 631 770 periods of microwave (MW) radiation stabilized to the ground state hyperfine transition in 133cesium [1]. The physical implementation is a global network of primary cesium frequency standards; Cs atomic fountain clocks have reached an uncertainty level of 4.5 × 10−16 [2]. In fountain clocks the interrogation time is limited by the ballistic flight time of the atoms within the apparatus; further progress within this approach can be achieved in space-based realizations such as the PHARAO project [3].

The recent progress in atom cooling and trapping together with the emergence of sub-hertz linewidth lasers allows development of more accurate frequency standards based on optical transitions using trapped neutral atoms and ions. For example, an inaccuracy level evaluation for a neutral 87Sr optical lattice clock is 1.5 × 10−16 [4], and the 27Al+ quantum logic frequency standard demonstrated an inaccuracy level of the order of 10−17 [5]. These approaches benefit from a 104–105 gain in quality factor by going from MW to optical transitions and longer observation times using trapped quantum oscillators.

A promising candidate for the role of a clock transition in a next-generation quantum frequency standard is the nuclear isomer transition in 229thorium. This unique radioisotope has an extremely low-energy 229mTh isomeric state. The most recent estimate of the transition energy is 7.8 ± 0.6 eV, obtained from the indirect measurements of the 233U → 229Th decay spectrum with an advanced x-ray microcalorimeter [6]. The corresponding wavelength of about 160 nm in the vacuum ultraviolet (VUV) is attainable by modern laser systems, and hence construction of a thorium 'nuclear clock' seems plausible.

An optical frequency standard based on a nuclear transition will open up new possibilities for experimental studies of basic laws of physics, i.e. constraining possible drifts of fundamental constants of interactions. In [7] it is conjectured that a possible drift of the fine structure constant δα/α will result in a corresponding but amplified change of the 229Th nuclear transition frequency δω/ω with an amplification factor between 102 and 104. Hence, the thorium quantum frequency standard will be several orders of magnitude more sensitive to possible drifts in α than existing standards based on electronic shell transitions in atoms or ions.

It should be noted that the current energy value of the thorium isomer transition is derived indirectly by looking at γ transitions with energies much higher than the isomer one. Therefore this value cannot be considered definitive until a direct measurement is made. Several groups worldwide (i.e. PTB, Germany; Georgia Institute of Technology, USA; University of California, Los Angeles, CA, USA; Vienna University of Technology, Austria) are currently working toward this goal.

Two main experimental approaches to a thorium clock, first proposed in [8], are under realization now: the use of trapped thorium ions [911] and 'solid-state nuclear spectroscopy' [12, 13].

In the first scheme, individual or small ensembles (up to 105) of ions are trapped and laser cooled using established techniques. The main advantage of the trapped ion approach is precise knowledge of and control over the fields affecting the ions. It has been proposed that a nuclear clock based on a single trapped 229Th3+ ion might reach a fractional inaccuracy of 10−19 [17].

In the second approach, a macroscopic number (up to 1018) of thorium atoms is implanted into a VUV-transparent crystal and interrogated simultaneously. This scheme will require none of the complex experimental techniques for ion trapping and cooling and can be fabricated in mass production. The number of simultaneously excited thorium nuclei is many orders of magnitude higher than in the trapped ion approach, therefore the solid-state spectroscopy method seems to be more promising for an initial localization of the frequency of the isomer transition.

In this work, we analyze the possibility of constructing a 'solid-state nuclear clock' based on 229thorium-doped crystals. Here, certain difficulties arise from the systematic crystal effects and the very long relaxation time of the excited nuclear state that can significantly exceed the interrogation time. We briefly review the current state of knowledge about the thorium isomer energy and lifetime and consider transition line shifts, splittings and broadening processes due to the crystal environment for the specific case of thorium doped into calcium fluoride (CaF2) crystals (section 2). We then describe the interaction of the nuclear quantum states of the thorium doping complex with coherent laser radiation in section 3. In section 4, we propose an interrogation scheme based on the counting of fluorescence photons to implement a solid-state nuclear clock and estimate the attainable fundamental limits of precision. At liquid nitrogen temperatures, we find that the clock performance is determined by spin–spin relaxation due to interactions of the thorium nuclear moment with neighboring nuclei. A detailed analysis of this process (presented in section 5) for the thorium-doped CaF2 system indicates that a fractional instability at the 10−19 level might be reached (neglecting technical limitations imposed by the interrogation laser system) within the solid-state clock approach.

2. Thorium nuclear transition in the crystal lattice environment

2.1. Transition energy and isomer state lifetime

The isomer transition energy of 229Th is currently derived from indirect measurements of the γ-ray spectrum resulting from the decay of 233uranium. In 1989–1993 Helmer and Reich performed the first measurements using high-quality germanium detectors (resolution from 300 to 900 eV). They estimated the 229Th isomer state energy to be E = 3.5 ± 1.0 eV [18, 19]. This unnaturally low value triggered a multitude of investigations, both theoretical and experimental, trying to determine the transition energy precisely and to specify other properties of the isomer state of 229Th (such as the lifetime and the magnetic moment). However, searches for direct photon emission from the low-lying excited state have failed to report an unambiguous signal [2023]. In 2005, Guimãraes-Filho and Helene re-analyzed the old experimental data and reported the value of E = 5.5 ± 1.0 eV [24]. New indirect measurements with an advanced x-ray microcalorimeter (resolution from 26 to 30 eV) were carried out by Beck et al in 2007 [25]. They published a new value for the transition energy E = 7.6 ± 0.5 eV for the isomer nuclear transition, shifting it into the VUV domain. This shift probably explains the absence of signatures of the transition in previous experiments. In 2009, Beck et al [6] reviewed their results taking into account the nonzero probability of the 42.43 keV → 229mTh transition (estimated as 2%) and published a revised version E = 7.8 ± 0.5 eV. This value is currently the most accepted one in the community but cannot be considered definite until a direct measurement is made successfully. The dominant uncertainty in the prediction of Beck et al [25] is connected to the value of the branching ratio b = 1/13 from the 29.19 keV level to the ground state [26]. Estimations of this value in various references differ. In [26], two alternative values for this branching ratio are mentioned: 25% that would result in E = 9.3 ± 0.6 eV and 51% that would result in E = 14.0 ± 1.0 eV. In the present analysis, we use the value of 7.8 ± 0.5 eV; scaling the obtained results to different transition energies is straightforward.

While a direct measurement of the isomer transition energy in 229Th is outstanding, the decay rate γ of the isomer state also remains unknown. The most recent estimations of the half-life of the isomer state for a bare nucleus are based on theoretical calculations of the matrix element of the magnetic moment [27]. The theory was verified by a comparison with experimental data for transitions at higher energies. It predicts a half-life of T1/2 = (10.95 h)/(0.025E3) for the isomer transition, where E is given in eV. For E = 7.8 eV, this yields T1/2 = 55 min, which corresponds to a spontaneous decay rate of γs = 2.09 × 10−4 s−1. In [28], it is predicted that the probability of spontaneous magnetic dipole emission in a transparent non-magnetic dielectric medium with refractive index n is n3 times higher than that in vacuum. The reason for this factor is the renormalization of the density of emitted photon states. For a CaF2 crystal, n = 1.55 at a wavelength λ = 160 nm (corresponding to E = 7.8 eV); the enlargement factor is 3.7 for the spontaneous decay rate of the isomeric state. Therefore, if the crystal contains Ne excited thorium nuclei, the number dNph of 160 nm photons emitted during a short time interval dt can be estimated as

Equation (1)

The total decay rate γ is determined not only by the spontaneous decay of the nucleus; it can be enhanced by interactions between the nucleus and the electron shell in electronic bridge processes and bound internal conversion effects [29, 31, 32]. Such a decay can be accompanied by the emission of several low-energy photons with widely distributed energies [29]. This would make it difficult to separate decay events from optical noise sources (i.e. from scintillation or radioluminescence).

There are a number of theoretical calculations on electronic bridge processes in the thorium system. In an isolated neutral atom the thorium isomer state lifetime is expected to be about 4.5 min for E = 3.5 eV and about 10−5 s for E = 7.6 eV [29]6 . In the Th3+ ion for E = 7.6 eV, the expected lifetime is the same as that in a bare nucleus described above if the valence electron is in the ground 5f5/2 state. It decreases by a factor of 20 if the valence electron is in the metastable 7s state [31]. In the Th+ ion the lifetime decreases 102–103 times for E = 3.5 and 5.5 eV [32]; there are not sufficient data about the Th+ ion electronic structure to calculate the lifetime at E = 7.8 eV. Similar calculations for the Th4+ ion, as will occur in the solid-state approach presented here, are yet to be performed.

There have been several experimental attempts to determine the 229mTh lifetime without knowledge of the exact isomer energy. Inamura and Haba [33] measured the temporal variation of the α decay activity in a sample of 229Th after excitation in a hollow-cathode electric discharge. They proposed a value of T1/2 = 2 ± 1 min for the half-life of the isomeric state, which corresponds to a total excited state decay rate of γ = 6 +6−2 × 10−3 s−1. Kikunaga et al [34] studied the α spectrum of chemically 'fresh' 229Th produced in the α decay of 233U. The authors obtain T1/2 < 2 h with 3σ confidence. The latter experiment was not suited to measuring half-lives of the order of minutes because of the long chemical preparation process. The analysis of both the experiments is significantly complicated by the presence of various chemical compositions of thorium, mainly ThO+, Th and Th+ in [33] and hydroxide and chloride complexes in [34]. These chemical structures determine the energies of electronic levels and hence the rates of electronic bridge processes. Therefore the different results obtained in [33, 34] may be explained by the different chemical structures and the times of preparation and measurement.

In this work, we assume the band gap of the thorium-doped CaF2 material to be sufficiently large to neglect electronic bridge processes and contributions from low-energy photons to the measured fluorescence count rates. In other words, we expect no electronic levels near 7.8 eV in the thorium-doped CaF2 crystal. This assumption will be substantiated in a future publication [16]. We perform our analysis using the theoretical value for the bare nucleus lifetime γ = γs·n3 = 7.8 × 10−4 s−1, enhanced by the refractive index, and assume the relaxation via electronic bridge processes to be negligible. Again, as for the energy, scaling our results to different values of the 229mTh lifetime is straightforward.

2.2. Effects of the crystal lattice environment

In this work, we consider the particular system of thorium nuclei doped into calcium fluoride single crystals. CaF2 is a standard material in UV optics; it is routinely grown and doped in many laboratories and companies worldwide using various techniques (Czochralski, Bridgman–Stockbarger and micro-pulling down). Standard procedures for cutting, cleaving, polishing and coating are established; CaF2 is well characterized concerning radiation damage, both concerning UV laser radiation and radiation related to the nuclear decay chain of 229Th.

CaF2 has a wide band gap of about 12 eV; the values in the literature range from 11.6 eV [14] to 12.1 eV [15]. The valence band of CaF2 consists of 2p levels of the fluorine ions, while the bottom of the conduction band originates from the 4s and 3d orbitals of the calcium ions. Optical transmission of better than 70% in the 160 nm wavelength regime is routinely achieved in high-purity single crystals.

The simple cubic lattice structure and the fact that only the Ca2+ site qualifies as a doping site make thorium-doped CaF2 amenable to full electronic structure simulations [16], which is currently out of reach for much more complex crystal structures (i.e. LiCaAlF6 as proposed in [12]).

Thorium atoms in a solid-state crystal lattice are confined in the Lamb–Dicke regime; the recoil energy Erec = (7.8 eV)2/(2MThc2) ≃ 1.4 × 10−10 eV is far below the energy required for creating a phonon with a recoil momentum prec = 7.8 eV/c in the lattice. Therefore the only process that could transfer the recoil energy to the lattice vibrations must include the scattering of optical phonons. However, optical modes are expected to be frozen out even at room temperature [12]. Therefore, internal and external degrees of freedom are decoupled and there is no sensitivity of the nuclear transition to recoil or first-order Doppler effects. However, the interaction between the thorium nuclear levels and the crystal lattice environment can cause some inhomogeneous effects which will affect the performance of a solid-state nuclear clock.

We assume the 229Th ions to be chemically bound into a VUV transparent ionic crystal. Therefore the fine interaction of the thorium nuclei with the crystal lattice environment is absent because of the absence of unpaired electronic spins. The remaining interaction is the hyperfine interaction, where the Hamiltonian $\skew3\hat {H}'=\skew3\hat {H}_{\mathrm {HFS}}$ can be represented in a multipole expansion $\skew3\hat {H}_{\mathrm {HFS}}=H_{E0}+H_{M1}+H_{E2}+\cdots $  [12]. In the following, we will discuss the individual terms in more detail.

The electric monopole shift (HE0) arises from contact interaction between the electron cloud and the finite nuclear volume. As the ground state and the excited nuclear state can have different sizes, this effect leads to a shift of the nuclear transition relative to a bare nucleus. Typical monopole shifts are about 1 GHz [35]; they will depend on the specific choice of the host crystal. However, this effect will be identical (up to temperature effects) for all nuclei having the same position within the lattice and hence deteriorate the solid-state nuclear clock accuracy, but not the stability. Temperature effects appear because the electron density at a given lattice site is temperature dependent. This temperature dependence is estimated as 10 kHz K−1 [12], which leads to a broadening of about 10 Hz for a temperature stability of ΔT ≈ 1 mK, that is attainable by standard techniques. It should be noted that a certain contribution to the decoherence rate can come from crystal lattice vibrations. These vibrations will modulate the electronic density in the vicinity of the nucleus and thus broaden the transition line. A detailed study of this broadening effect is beyond the scope of this paper. We expect that this effect, if relevant at all, can be effectively suppressed by cooling the crystal lattice.

The electric quadrupole shift (HE2) appears due to the electric quadrupole moment Q of the 229Th nucleus in the presence of a non-vanishing electric field gradient at the position of the nucleus. Reported experimental values for the quadrupole moment in the ground state are Qg = 3.149(32) eb [36] and Qg = 3.11(16) eb [37]; the value Qis of the quadrupole moment in the isomer state is still unknown; estimations [38] give Qis ≈ 1.8 eb. If the electric field gradient in the principal axes has two equal components, ϕxx = ϕyy = −ϕzz/2, the matrix element of the quadrupole interaction term can be written as [39]

Equation (2)

where the z-axis is the axis of symmetry of the electric field, ϕzz = ∂2φ/∂z2 and I and m are the nuclear angular moment and its projection on the axis of symmetry of the electric field, respectively. In a Th-doped CaF2 crystal, the thorium ion will most probably replace a Ca2+ ion in a Th4+ state. To recover an ionic (and hence transparent) crystal, the two remaining electrons will be localized by a charge compensation mechanism, either by introducing two additional fluorine (F) interstitials or by creating a vacancy by leaving a neighboring Ca2+ site empty. First simulations indicate a preference for charge compensation by interstitials which will be the main contribution to the electric field gradient [40].

Consider the charge compensation by two F interstitials as shown in figure 1(b). We expect the geometrical lattice distortion around the dopant to be negligible. Using the Ca2+ lattice constant a = 5.6 Å and the quadrupole antishielding factor γ = −177.5 [41], we estimate ϕzz ≈ −5.1 × 1018 V cm−2. This leads to quadrupole splittings of the order of a few hundreds of MHz, which can be easily resolved in laser spectroscopy (see figure 2). Note that this splitting could also be directly measured experimentally using the standard NMR techniques, which would give valuable information on the details of the charge compensation mechanism.

Figure 1.

Figure 1. CaF2 crystal lattice (a) and thorium inclusion with charge compensation by two in-line fluorine interstitials (b).

Standard image
Figure 2.

Figure 2. Quadrupole structure of the ground and isomer state of the 229Th nucleus in the CaF2 crystal lattice with charge compensation as presented in figure 1(b). Light arrows represent the possible magnetic dipole transitions, matrix elements are proportional to the indicated Clebsch–Gordan coefficients CI'm'Im1q (q = m' − m). The quantization axis z is chosen along the axis of symmetry of the electric field gradient.

Standard image

The quadrupole interaction of the nucleus with phononic modes of the crystal lattice leads to inhomogeneous broadening and mixing between different quadrupole sublevels as the typical distances between the quadrupole sublevels are much shorter than characteristic phonon energies. Comparison with experimental data for the quadrupole spin-lattice relaxation of the halogen nuclei in CsBr and CsI crystals [42] and scaling the theory of Van Kranendonk [43] and Yoshida and Moriya [44] to the CaF2 crystal yield a relaxation and mixing rate of the order of 2π·(5–100) Hz at room temperature. This rate scales as T2, leading to about 2π·(0.3–7) Hz at liquid nitrogen temperature. The temperature of the crystal should, however, not be so low as to entirely suppress the mixing, as populations accumulating in quadrupole sublevels not interacting with the laser field could lead to a loss of signal or necessitate additional repump lasers.

The second-order Doppler effect leads to a shift and broadening of the isomer transition due to vibrations of the nuclei. Using a classical harmonic oscillator model, we find the shift S = −ωv2〉/(2c2) = −3ωkBT/(2MThc2) and the broadening $\Delta \omega =\omega \langle \Delta v^2 \rangle /(2 c^2) = \sqrt {6}\,\omega k_{\mathrm {B}} T/(2M_{\mathrm {Th}}c^2)$ . Here ω is the isomer transition frequency, v the velocity of the thorium nuclei, $\langle \Delta v^2 \rangle =\sqrt {\langle v^4 \rangle - \langle v^2 \rangle ^2}$ . This leads to a shift of about 2π·1.14·T  (in Hz) and to a broadening of about 2π·0.93·T  (in Hz) where T is the temperature in kelvins. At liquid nitrogen temperature, we expect a shift of about 2π·88 Hz and a broadening of about 2π·72 Hz.

The term HM1, i.e. the magnetic dipole interaction with randomly oriented spins of surrounding nuclei, also contributes to broadening. The characteristic energies of spin–spin interactions correspond to the temperature range of the order of 1 μK; the broadening is expected to be temperature independent for experimentally relevant conditions, in contrast to all other broadenings discussed above, which decrease with decreasing temperature. A detailed evaluation (see appendix A) yields a contribution to the optical relaxation rate from ∼2π·84 Hz for the $|I=5/2,m={\pm } 3/2\rangle \leftrightarrow |I'=3/2,m'={\pm } 1/2\rangle $ optical transition to ∼2π·251 Hz for the $|I=5/2,m={\pm } 5/2\rangle \leftrightarrow |I'=3/2,m'={\pm } 3/2\rangle $ optical transition.

To proceed, we expect a global uniform shift of about 1 GHz dominated by the electric monopole shift compared to the bare nucleus isomer transition frequency (assuming identical doping complexes and charge compensation). The main sources of inhomogeneous broadening are magnetic dipole interaction and the second-order Doppler effect, leading to an ensemble linewidth of 2π·(84–251) Hz and 2π·0.93 Hz · T, respectively, where T is the temperature in kelvins. Taking the ensemble linewidth of 2π·150 Hz, we expect a quality factor Q = 1.26 × 1013 for the solid state clock transition. Note that crystal field shifts and broadenings (and their temperature dependence) can be directly studied with narrow band lasers, transferring the Mößbauer spectroscopy into the optical regime.

3. Excitation of thorium nuclei by laser radiation

Let us consider an ensemble of 229Th4+ ions implanted into a CaF2 crystal and interacting with a narrow-band VUV laser whose frequency ωL is close to the frequency ω of the isomer transition. Here we consider only the ions in one charge-compensating configuration, i.e. the one presented in figure 1(b). All other configurations will not interact with the laser field due to different quadrupole shifts. Th4+ ions are iso-electronic with the noble gas radon with zero angular momentum and zero spin due to the electronic shell. In the following, we neglect the electronic shell structure as well as higher excitation levels of the thorium nucleus. The nuclear ground and isomer states are degenerate with respect to the sign of projections m and m' and with respect to angular momenta I and I' up to interactions with randomly oriented spins of surrounding nuclei. The magnetic field $\skew5\vec {B}$ of the spectroscopy laser radiation can be written as $\skew5\vec {B}=\skew5\vec {B}_0 \, \cos \omega _{\mathrm {L}} t$ .

The Hamiltonian $\skew3\hat {H}$ of the thorium nucleus in resonance approximation can be represented as

Equation (3)

Here, the indices e and g indicate the sublevels of the nuclear excited and ground states, respectively, Veg = B0qμqeg/(2ℏ), B0q is the qth covariant component of $\skew5\vec {B}_0$ , μqeg is the qth contravariant component of the matrix element $\skew5\vec {\mu }_{eg}$ of the nucleus magnetic momentum operator in cyclic coordinates, q = m' − m, and $\skew3\hat {H}_0$ is the Hamiltonian of the free thorium ion. The term $\skew3\hat {H}'$ describes the interactions of the nucleus with the crystal lattice leading to shifts (which could also be added to $\skew3\hat {H}_0$ ) and relaxation terms that will be discussed later. The set of equations for the density matrix ρij with relaxation terms in the rotating frame reads as

Equation (4)

Equation (5)

Equation (6)

Equation (7)

Equation (8)

Here Γe,g and Γi,j|| are decoherence and mixing rates between magnetic states of the same quadrupole energy level, Γ' is the relaxation rate of optical coherence ρeg, ωij is the difference between the frequencies of the ith and jth levels and γeg = γ·|CI'm'Im1q|2 is the rate of spontaneous |e〉 → |g〉 transitions.

In this work, we suppose that the spectroscopy laser field is weak, i.e.

Equation (9)

Then, from equations (4)–(8) and from the normalization condition $\sum _i \rho _{ii}=1$ , we derive:

Equation (10)

where $\rho _{\mathrm {exc}}=\sum _{e}\rho _{ee}$ is the total excited state population. From equations (6), (8) and (10), we obtain

Equation (11)

The quadrupole splitting in the ground and the isomer state is of the order of hundreds of MHz, which is much greater than the expected value of Γ' (about a few kHz). Hence, only transitions between certain quadrupole sublevels of the ground and excited states should be taken into account. Let us consider the case when the laser is tuned close to the transition between |g1,2〉 = |I = 5/2,m = ± 3/2〉 and |e1,2〉 = |I' = 3/2,m = ± 1/2〉. This is the transition with the lowest relaxation rate of optical coherence; see appendix A. In the following, we will name this transition the clock transition and use ωeg = ω. We introduce the optical detuning Δ = ωL − ω. Suppose, for the sake of simplicity, that the radiation has a certain polarization along the q'th cyclic unit vector, so $B_{0q}=|\skew5\vec {B}_{0}|\delta _{qq'}$ . In dipole approximation, $\gamma _{\mathrm {s}} = \frac {4 \omega ^3}{3 \hbar c^3}\sum _g|\mu ^q_{eg}|^2$ . According to the Wigner–Eckart theorem, $\mu _{eg}^q=\Vert\mu \Vert C_{Im1q}^{I'm'}/\sqrt {2I'+1}$ , where $\Vert \mu \Vert=\sqrt {\frac {4\pi }{3}B(M1)(2I'+1)}=0.65 \, \mu _N$ is a reduced dipole moment (B(M1) = 0.025 μ2N is the reduced transition probability [27]), CI'm'Im1q is a Clebsch–Gordan coefficient and c is the speed of light in vacuum. Using $\sum _{m,q}\left |C_{Im1q}^{I'm'}\right |^2=1$ , we express

Equation (12)

and therefore

Equation (13)

Averaging over possible spatial orientations of the electric field gradient with respect to the polarization of the laser, and expressing the amplitude $|\skew5\vec {B}_0|$ via the intensity I0 of the resonance radiation: $|\skew5\vec {B}_0|^2=8 \pi I_0/c$ , we finally obtain

Equation (14)

where the excitation rate R is

Equation (15)

Here we denote the averaging over spatial orientations by angled brackets and we have taken into account the values of the 229thorium nuclear angular momenta I = 5/2 and I' = 3/2 in the ground and the excited state, respectively.

4. The interrogation scheme and fundamental limits of stability

Any passive quantum frequency standard consists of three main elements: a generator producing some periodic signal (a laser in the case of an optical frequency standard) with good short-time stability and a frequency that can be tuned around a 'nominal' frequency ωn, a quantum discriminator, and a servo loop determining and correcting the offset f = ωn − ω between the nominal frequency ωn and the frequency ω of the clock transition according to the result of the interrogation of the quantum discriminator. If we add a system to count the number of oscillations of the generator, we obtain a clock (see figure 3).

Figure 3.

Figure 3. Schematic representation of a passive quantum frequency standard constructed from a generator (laser), a quantum discriminator and a feedback loop. Linking the stabilized generator to an electronic counting system (i.e. through a frequency comb system) allows the realization of a clock.

Standard image

Usually, the Rabi, Ramsey [45] or hyper-Ramsey [46] interrogation scheme is used in high-performance frequency standards. However, all these methods require exact knowledge of the quantum state of the quantum discriminator at the beginning of any interrogation cycle. If the quantum discriminator consists of one or a small number of ions it is possible to know the quantum state precisely at the end of any interrogation cycle (and therefore at the beginning of any next cycle) as the interrogation measurement leads to a projection of the quantum state onto a well-defined basis. If the quantum discriminator consists of a large number of trapped ions it is possible either to reload the trap with ions of a known quantum state or to repump the ions back to the initial quantum state via some short-living state. Note that for thorium ions in an ion trap such a 'repumping' of the isomer state could be performed using an excitation of the electronic shell to a resonant metastable state which will lead to an amplification of electronic bridge processes [31].

In the solid-state nuclear frequency standard, such a repumping seems to be impossible due to the absence of an accessible metastable state in the band gap. Another fundamental problem arose from the huge difference between the relaxation time of the excited state population (presumably tens of minutes) and the short coherence time between the ground and excited states (milliseconds) due to crystal lattice effects. At least one Rabi or Ramsey interrogation cycle should be completed during the time interval t < 1/Γ', whereas the time tp needed to bring the nuclei back to the ground state should be tp > 1/γ. Therefore the Rabi or Ramsey interrogation time would take only a tiny (≈10−6) fraction of the total interrogation cycle, which would significantly reduce the clock performance due to the Dick effect [47]. Moreover, Ramsey or Rabi interrogation schemes require quite strong laser fields, able to excite a significant number of nuclei in a short interrogation time.

4.1. Clock interrogation by fluorescence spectroscopy

To tackle these problems, we propose an interrogation scheme based on counting the spontaneous nuclear decay fluorescence photons after illuminating the quantum discriminator with the spectroscopy laser. The (n, n + 1) interrogation cycle consists of four time intervals, as depicted in figure 4. In the first interval θ between tn and t'n, the crystal sample is illuminated by laser radiation with the frequency detuned by δm to the blue side of the nominal frequency ωn of the generator. Then a laser shutter is closed and the fluorescence photons are counted by a photodetector during the second time interval θ' between t'n and tn+1. Then these two operations are repeated, but with the laser frequency shifted to the red side by δm of the nominal value. After the last measurement, we determine the frequency offset f = ωn − ω and make a correction to the nominal frequency. As no further state initialization, repumping or sample preparation is required, the next interrogation cycle can start immediately.

Figure 4.

Figure 4. Interrogation cycle of the quantum discriminator as discussed in the text. Note that this cycle is operated continuously without any dead time.

Standard image

We will now consider a single interrogation cycle in more detail. Solving equation (14), we express the excited state population ρexc at time t'n as

Equation (16)

The number Nn of nuclear fluorescence photons counted by the photodetector during time θ' between t'n and tn+1 is

Equation (17)

where

Equation (18)

is the product of the number NTh of interrogated thorium nuclei, the ratio γsn3/γ of the rate of radiative decay of the isomer state to the total decay rate enhanced by possible electronic bridge processes, and the probability of detecting the emitted photon. Here k is the quantum efficiency of the photodetector and Ω is the effective solid angle covered by the detector.

In turn, the excited state population at time tn+1 is equal to

Equation (19)

and at time t'n+1 it is equal to

Equation (20)

and in turn

Equation (21)

Let us introduce the 'evolution rate in the presence of laser field'

Equation (22)

and the functions

Equation (23)

Equation (24)

We then can rewrite expressions (17) and (21) as

Equation (25)

Equation (26)

where Nn−1 is the number of photon counts collected in the last measurement of the previous interrogation cycle, i.e. between t'n−1 and tn.

We suppose that the generator offset f is small, i.e. f ≪ Γ'. We can then expand the functions a and b near the working point δm into a Taylor series:

Equation (27)

Using equations (25)–(27), we express the frequency offset f as

Equation (28)

Now let us calculate the error δf in determining the frequency offset f. Here we suppose that Nn+1 and Nn are Poissonian random numbers with mean values described by (25) and (26), i.e. the error of Ni is $\sqrt {\bar{N}_i}$ . The average value $\bar{N_{i}}$ is

Equation (29)

From (25)–(29), we express the error of the frequency offset as

Equation (30)

Substituting explicit expressions for a0, b0, a1 and b1 into (30), we obtain

Equation (31)

where G = G(δm).

Expression (31) represents a fundamental lower limit on the error of the frequency offset for one interrogation cycle as depicted in figure 4. If the generator has good long-term stability, and the algorithm for frequency stabilization succeeds in the total correction of the measured offset (28), then expression (31) is the frequency error corresponding to an absolute frequency instability of the solid-state nuclear clock after one interrogation cycle.

The frequency error (31) depends on the specific properties of the quantum discriminator (NTh, γ and Γ'), on the detection system (k Ω), on the intensity of laser radiation determining the excitation rate R and on the regime of interrogation (θ, θ' and δm). In practice, the total interrogation time t = 2(θ + θ') will be imposed by the short-term stability of the interrogation laser (i.e. the duration over which frequency drifts and fluctuations are smaller than δf). Also the laser intensity (and therefore the excitation rate R) will have a technical upper limit. On the other hand, the working point δm and the relative irradiation time

Equation (32)

can be adjusted freely. It is therefore of interest to determine the optimal values of tR and δm and study the dependence of the fractional frequency error δf/ω as a function of total interrogation time t and excitation rate R for optimized tR and δm. Because the general expression (31) seems to be too complex for a direct interpretation, we will first focus on the limiting cases of short and long interrogation times without restricting ourselves to explicit experimental parameters. In section 5, we will present numerically optimized results for the specific system of thorium-doped calcium fluoride crystals. A comparison of the analytic expressions with numerical optimizations can be found in appendix B.

4.2. Frequency error for short interrogation time

Short interrogation time t = 2(θ + θ') refers to γθ' ≪ 1 and Gθ ≪ 1, so the irradiation time and the detection time are shorter than the overall decay time and the inverse excitation rate, respectively. In this case we can expand the exponents in (31) into a Taylor series and keep only the main terms. We obtain

Equation (33)

and the absolute frequency error scales as t−3/2 with the total duration of the interrogation cycle. First we fix δm and t and find the optimal value of tR. Therefore we should minimize the function

producing the optimal value for tR:

Equation (34)

The frequency error δf for the optimized value of tR is

Equation (35)

Taking into account (22) and minimizing (35), we can find optimal values for the working point δm and the frequency error δf.

In the limit of a weak field, i.e. for R ≪ γ, we find that

Equation (36)

Equation (37)

In the limit of a strong field, i.e. if R ≫ γ, we find that

Equation (38)

Equation (39)

4.3. Frequency error for long interrogation time

We now consider the case when the interrogation time is long compared to the total decay time, i.e. when γt ≫ 1. Here we cannot neglect the exponents in (31) because the optimal value either of θ or of θ' can be comparable to 1/γ or 1/G, respectively. From the structure of (31) we expect a small optimal value of θ because of the presence of $\sqrt {1-\mathrm {e}^{-G\theta }}$ in the numerator. Therefore, we suppose that γθ' ≫ 1 and neglect the terms in eγθ'. Under these assumptions the frequency error of the nuclear clock can be expressed as

Equation (40)

where x = Gθ and α = γ/G. To find the optimal value for θ, we minimize (40) with respect to x, which leads to the transcendental equation for x:

Equation (41)

This equation yields the optimal value of θ = x/G for the case of a long total interrogation time. We again consider the cases of weak and strong interrogation fields.

If the laser field is weak, i.e. if R ≪ γ, α is close to 1. Then 1 − α = 5R/(2γ(1 + δ2m/Γ'2)) and (41) transforms into

This equation contains a large term proportional to γ/R on the right side of the denominator which has to be compensated for by another large term 2x0 − 1 to yield a solution x0 of (41). We obtain

Equation (42)

Therefore, for the case of a weak laser field, Gθ ∼ x0 ≫ 1, and we can neglect eGθ in (40). Optimization for δm yields

Equation (43)

Equation (44)

We now consider the case when the laser field is strong, i.e. when R ≫ γ. We then can set α = 0 in (41). We obtain

Equation (45)

In this case, the optimal working point δm and the frequency fluctuation δf are

Equation (46)

Equation (47)

The last expression determines the ultimate stability limit that can, in principle, be attained by the fluorescence spectroscopy method in a single interrogation cycle for a given quantum discriminator. It is valid for a large range of parameters as discussed in appendix B. The optimized value for δf (47) is exclusively determined by the decoherence rate Γ' (which is estimated in appendix A) and the number of interrogated nuclei; it does not depend on the exact values of the isomer transition energy or lifetime. To reach this performance level, the conditions R ≪ Γ' and R ≪ Γgg' have to be fulfilled; otherwise an accumulation of populations in quantum states not interacting with the laser field will decrease Neff over time. However, the very large ratio of Γ'/γ ∼ 106 gives a broad range of parameters where R ≫ γ and R ≪ Γ',Γgg' can be satisfied simultaneously.

Note that in the limit of long interrogation time, the equilibrium value of the excited state population is attained in a short fraction of the interrogation cycle and the optimal excitation time takes only a small fraction of the interrogation cycle. In contrast, in the limit of short interrogation time, the equilibrium value cannot be attained in one interrogation cycle and the optimal excitation time θ should be equal to the photon collection time θ'.

5. Performance of a nuclear clock based on thorium-doped calcium fluoride

In this section, we present numerically optimized relative irradiation times tR, ideal working points δm and hence ultimate frequency performances of the solid-state nuclear clock scheme for different excitation rates R and interrogation times t. We use the specific system parameters for 229Th ions doped into CaF2 crystals introduced in the previous sections: γ = γsn3 = 7.8 × 10−4 s−1, Γ' = 150 Hz·2π = 942.5 s−1, and ω = 2π·1900 THz = 1.19 × 1016 s−1. We suppose that the total number of thorium nuclei is NTh = 1014, the solid angle is Ω = 4π/10 and the quantum efficiency of the photodetector is k = 0.1 (which is typical of photomultipliers in our wavelength range). We obtain Neff = 1012, see (18).

The optimized parameters δm and tR are represented in figure 5 for different excitation rates R. One can see that for all interrogation times, the optimal working points δm lie between Γ' and $\Gamma '/\sqrt {2}$ , which are the limiting results for R ≪ γ ((36), (43)) and R ≫ γ ((38), (46)), respectively.

Figure 5.

Figure 5. Dependence of the optimal working point δm (a) and of tR (b) on the total interrogation time t for different values of the excitation rate R.

Standard image

In general, the ideal working point is close to the inverse decoherence rate of the order of kilohertz (see appendix A). This means that the linewidth of the interrogation laser has to be significantly narrower than this; otherwise feedback stabilization to the nuclear transition will fail. Also one can see that the optimal value of tR decreases with decreasing t and increasing R, which means a shorter excitation than detection time. This is connected to the fact that the optimal illumination time θ should not significantly exceed 1/G(R,δm), i.e. the typical time scale to approach the equilibrium population of the excited state $\rho _{\mathrm {exc}}^{\mathrm {eq}}=\frac {2}{5}(1-\gamma /G(R,\delta _m))$ according to (14).

Let us now turn to the nuclear clock error per interrogation cycle obtained for the optimized parameters δm and tR, expressed as fractional frequency error in figure 6. For short interrogation times, i.e. t < 0.1 γ−1, the fractional frequency error scales as δf/ωt−3/2 in accordance with (37) and (39). For long interrogation times, i.e. t > 10 γ−1, the frequency error δf/ω becomes independent of the interrogation time in accordance with (44) and (47). We also note that for strong pumping, i.e. R > 30 γ, the clock performance is practically independent of the exact value of R; for long interrogation times this is already the case for R > γ. For t > 10/γ and R more than a few γ, the fractional frequency error δf/ω per interrogation cycle is essentially independent of R and t and is equal to the ultimate value of about 1.8 × 10−19 in accord with (47).

Figure 6.

Figure 6. Fractional frequency error δf/ω of the nuclear clock for different excitation rates R and different total interrogation times t.

Standard image

We now consider the parameter $\delta f/\omega \cdot \sqrt {t}$ which describes the fractional clock instability. It should be optimized with respect to interrogation time t if one wants to make a precise measurement of the frequency of the isomer transition (in the solid-state environment) by a long-term series of interrogations. It is also the relevant quantity describing the instability of the solid-state nuclear frequency standard when a narrow band laser is feedback-stabilized to the transition and compared to a primary frequency standard. If the series of measurements take a time τ = N·t, then the fractional instability σy(τ) of the series is equal to the fractional frequency error averaged over the series. In the shot noise limit we obtain

Equation (48)

Hence, to reduce σy(τ) one should optimize (minimize) $\delta f\sqrt {t}/\omega =\sigma _y\sqrt {\tau }$ . We find that the optimal value for t is of the order of a few γ−1, only weakly depending on the excitation rate R; see figure 7(a). The ultimate clock instability is given by the minimum value of $\sigma _y\sqrt {\tau }$ at the optimal value of t. It saturates at ≈1.63 × 10−17 s1/2 for high excitation rates (R/γ ≳ 10) and scales as R−1/2 for lower excitation rates (R/γ ≲ 0.1); see figure 7(b). The resonance intensity I0 of the laser source required for attaining the 'high excitation rate regime' R/γ ≳ 10 can be estimated using (15). For the system parameters used in this section it yields I0 ≃ 16 mW cm−2. Other optimized parameters corresponding to R = 10γ are t ≃ 3.88/γ ≃ 5 × 103 s, tR ≃ 0.026 and δm ≃ 0.99Γ'.

Figure 7.

Figure 7. Nuclear clock fractional instability $\sigma _y \sqrt {\tau }=\delta f/\omega \cdot \sqrt {t}$ as a function of the total interrogation time t for different excitation rates R (a) and the ultimate clock instability $\sigma _y^{\mathrm {opt}} \sqrt {\tau }$ obtained for the optimal value of t, as a function of R (b).

Standard image

A few words on the validity of approximation (9) and the optimal value of the temperature: taking R = 10γ and using (15), we easily estimate the 'reasonable' value of Veg as Veg ≃ 2π·0.53 Hz. The mixing rate Γgg' should be large in comparison with this value. Our estimation (see section 2.2) gives Γgg' ≃ 2π·(5–100) Hz for room temperature, and Γgg'T2; then we can take Γgg'[Hz] ≃ (5 × 10−5–10−3T2 [K]. Condition (9) gives $T[K] > \sqrt {\frac {0.53}{5\cdot 10^{-5}-10^{-3}}}$ or T > 20–100 K. If (9) is not fulfilled, in a real multilevel system a significant fraction of nuclei will be pumped into the dark state, and growth of the excitation rate R will reduce the performance. On the other hand, even at liquid nitrogen temperatures the contribution of the temperature-dependent second-order Doppler broadening term to the total decoherence rate Γ' becomes comparable to the temperature-independent magnetic dipole interaction term, and increasing the temperature is undesirable. We expect that the optimal sample temperature should be around the liquid nitrogen temperature. A more detailed analysis requires more precise data on the quadrupole relaxation and is beyond the scope of this paper. Nevertheless, we can claim that the optimal excitation rate R should be of the order of a few γ.

In an experimental implementation, the performance of the solid-state nuclear clock will be limited by technical constraints such as the temperature stability of the sample, the frequency and intensity stability of the interrogation laser system, and noise of the detection system. If these limitations can be overcome, the solid-state approach can reach the 10−19 fractional instability level within a few hours of averaging. It may hence be competitive with the trapped-ion scheme [17] concerning stability for a (temperature-stabilized) liquid nitrogen temperature sample. Note that the fluorescence stabilization scheme developed here is operated continuously without dead time between cycles as no state initialization or re-trapping of particles is required. The global line shifts (several hundreds of MHz) due to crystal fields will obviously degrade the clock accuracy and require calibration with a primary standard.

6. Conclusion

We have analyzed the possibility of constructing a 'solid-state nuclear clock' from 229Th nuclei implanted into VUV transparent calcium fluoride crystals. We expect static global line shifts to be dominated by the temperature-dependent electric monopole and electric quadrupole interactions and by second-order Doppler shifts, necessitating active temperature stabilization of the sample. The electric quadrupole interaction leads to a splitting of the nuclear ground and isomer levels of the order of a few hundreds of MHz and is the main source of state mixing. As broadening effects we considered again the second-order Doppler shift, the electric quadrupole and the magnetic dipole interaction. The magnetic dipole interaction does not depend on temperature and is hence expected to be the dominant source of decoherence (followed by the second-order Doppler effect) at liquid nitrogen temperatures. For the CaF2 crystal we estimate a decoherence rate of 150 Hz. The extreme gap between this rate and the nuclear state lifetime renders the application of commonly used coherent interrogation schemes (i.e. Rabi and Ramsey) inappropriate. We therefore developed an alternative clock stabilization scheme based on fluorescence spectroscopy. A detailed analysis reveals a favorable scaling with the number of interrogated nuclei, validating the solid-state approach. Neglecting all possible technical constraints (i.e. frequency and intensity stability of the interrogation laser), supposing that the sample contains 1014 thorium nuclei and that the detection probability is 1%, we find an ultimate fractional instability limit of 1.8 × 10−19 for a single interrogation cycle taking about 104 s.

Acknowledgments

This research was supported by the Austrian Science Fund (FWF) through projects M1272-N16 (TheoNAC) and Y481-N16; by the European Research Council (ERC) (starting grant 258603 NAC); by the State Foundation of Fundamental Researches of Ukraine (project F40.2/039); by the Russian Federal Program 'Scientific and scientific-pedagogical personnel of innovative Russia in 2009–2013'; by the Russian Foundation for Basic Research through grant RFBR-11-02-90426_Ukr_f_a; and by a grant from the President of the Russian Federation for young candidates of science (project MK-5318.2010.2). We thank P Mohn, P Dessovic and R Jackson for helpful discussion and calculations concerning the microscopic structure of the thorium-doping complex.

Appendix A.: Decoherence and mixing due to magnetic dipole interaction

A calcium fluoride crystal is formed from 19F fluorine and a natural mixture of calcium isotopes (40Ca, 42Ca, 43Ca, 44Ca and 46Ca). The only Ca isotope with nonzero nuclear magnetic moment, 43Ca, has a small abundance of the order of 0.14%, whereas the nucleus of the only abundant 19F isotope has a finite magnetic moment μF = 2.63 μN and a nuclear spin s = 1/2. In the following we consider only the magnetic moment of the 19F nuclei in the CaF2 lattice to interact with the thorium nuclear transition. The relaxation of the nuclear magnetic moment of thorium (or fluorine) nuclei is caused by the interaction with a random magnetic field $\mathcal {\skew5\vec {H}}$ at the position of the considered nucleus, created by the surrounding fluorine nuclei.

For the calculation of the relaxation and mixing rates, we use the formalism of relaxation theory presented in [48]. Here, we briefly describe this formalism and apply it first to the magnetic moment of fluorine nuclei. This yields the characteristic rate of fluorine spin–spin relaxation and therefore the characteristic correlation time of $\mathcal {\skew5\vec {H}}$ . Then we calculate the relaxation rate for the ultraviolet transitions between different quadrupole sublevels of the ground and isomer states of the thorium nucleus.

A.1. Relaxation method

Here we have followed a method developed in [48] for the calculation of the effect of a stationary random perturbation on the elements of the density matrix. The evolution of the density matrix is described by the Liouville–von Neumann equation

Equation (A.1)

Here $[\skew3\hat {A},\skew3\hat {B}]\equiv \skew3\hat {A}\skew3\hat {B}-\skew3\hat {B}\skew3\hat {A}$ is a commutator, $\skew3\hat {H}_0=\hbar \sum _\alpha |\alpha \rangle \,\omega _{\alpha }\langle \alpha |$ is the Hamiltonian of the unperturbed system, $\skew3\hat {H}_1^C$ is the Hamiltonian of interaction with the coherent external field and $\skew3\hat {H}_1$ is the Hamiltonian of interaction with a random field responsible for the relaxation, in our case with the random magnetic field $\mathcal {\skew5\vec {H}}$ . In order to eliminate the strong contribution $\skew3\hat {H}_0$ from (A.1), one applies the unitary transformation $\skew3\hat {L} \rightarrow \skew3\hat {L}'= \mathrm {e}^{\mathrm {i}\skew3\hat {H}_0t/\hbar }\skew3\hat {L}(\mathrm {e})^{-\mathrm {i}\skew3\hat {H}_0t/\hbar }$ to all the operators in (A.1) (interaction representation). We obtain the equation for the density matrix,

Equation (A.2)

The solution to this equation can be written as

Equation (A.3)

Substituting (A.3) into the second term on the right-hand side of (A.2) yields

Equation (A.4)

We assume that the characteristic time scale of the correlations of fluctuation of the random field $\skew3\hat {H}_1^{\prime }(t)$ is much smaller than all other characteristic time scales, including the characteristic time scale of variation of the density matrix7. Then one can choose the origin in such a way that the density matrix would not change noticeably during this time interval (one may substitute ρ'(t') by ρ'(t)) but all the correlations of the random field will disappear (one may take − as a lower limit of integration). Also suppose that all the matrix elements of $\skew3\hat {H}_1^{\prime }(t)$ have zero mean values (otherwise we would add them to $\skew3\hat {H}_1^{C\prime }$ ). Performing the averaging over the ensemble (indicated by overline), we obtain

Equation (A.5)

The first term in (A.5) corresponds to the evolution of the density matrix under the action of a coherent field. The second term describes the relaxation. If the perturbations are stationary, it can be written as

Equation (A.6)

Expanding the commutators in (A.6), one obtains

Equation (A.7)

where ωij = ωi − ωj,

Equation (A.8)

and

Equation (A.9)

(note that on the right-hand side of the last expressions, the matrix elements of $\skew3\hat {H}_1$ are taken in the 'initial' Schrödinger representation). Expression (A.7) for the relaxation operator $\mathbb {R}$ contains oscillating terms that can be neglected. As a result we keep only the terms where ωαβ = ωα'β'. Except for some special cases, this condition is fulfilled only if α = α' and β = β' (mixing of populations) or if α = β and α' = β' (relaxation of coherences).

Using (A.6)–(A.9) one can calculate the relaxation and population mixing rates.

A.2. Spin relaxation of a fluorine nucleus

Let us consider a single fluorine nucleus. The spin of this nucleus can take one of two projections m = ± 1/2 with respect to the arbitrary quantization axis z. Let us denote |1〉 = |m = −1/2〉 and |2〉 = |m = 1/2〉. The magnetic moment operator $\skew3\hat {\skew5\vec {\mu }}$ for the spin of this nucleus can be expressed as $\skew3\hat {\skew5\vec {\mu }}=\skew3\hat {\skew5\vec {s}}\, \mu _{\mathrm {F}}/s$ . Therefore, we can express the operator $\skew3\hat {H}_1=-\skew3\hat {\skew5\vec {\mu }} \cdot \skew5\vec {\mathcal {H}}$ as

Equation (A.10)

In this work, we use the 'diffusion' model for the correlations of the random magnetic field:

Equation (A.11)

The average square of the random magnetic field $\mathcal {H}_0^2$ and the correlation time tc will be determined later.

According to (A.8) the relaxation rate R1212 of the coherence ρ12 is

Equation (A.12)

From (A.9) and (A.11), one finds that

Equation (A.13)

where

Equation (A.14)

which yields the coherence relaxation rate

Equation (A.15)

Similarly, one finds that the population mixing rate is

Equation (A.16)

Now we can estimate the correlation time tc which is determined by the population relaxation rate:

Equation (A.17)

To find $\mathcal {H}_0^2$ we recall that in CaF2 the fluorine nuclei form a simple cubic lattice with a distance of a/2 = 2.73 Å between two neighbors. Spins of different fluorine nuclei are not correlated; therefore the average square $\mathcal {H}_0^2$ of the random magnetic field $\skew5\vec {\mathcal {H}}$ is the sum of squares of magnetic fields produced by different fluorine nuclei:

Equation (A.18)

$\skew5\vec {\mathcal {H}}_a$ is the magnetic field created by the ath fluorine nucleus whose radius-vector is $\skew5\vec {r}_a$ with respect to the nucleus of interest. Using $\skew3\hat {\skew5\vec {\mu }}_a=2 \skew3\hat {\skew5\vec {s}}_a\, \mu _{\mathrm {F}}$ and $\mathcal {\skew5\vec {H}}_a=(3\skew5\vec {r} (\skew3\hat {\skew5\vec {\mu }}_a\skew5\vec {r}_a) - \skew3\hat {\skew5\vec {\mu }}_a r_a^2)/r_a^5$ , we find that

Equation (A.19)

Therefore

Equation (A.20)

(here the sum is taken over l,m,n excluding l = m = n = 0) or $\mathcal {H}_{0}^2\simeq 21.5$ G2, which leads to a relaxation time tc ≃ 15 μs according to (A.17). Note that this value is quite close to experimentally observed spin–spin relaxation times ∼20 μs [49].

A.3. Spin relaxation of a thorium nucleus

We now consider an individual thorium nucleus interacting with the random magnetic fields created by the surrounding fluorine nuclei. The quantization axis is fixed by the direction of the main axis of the electric field gradient. Let us denote the different states of the thorium nucleus by |i〉, where i is a number ranging from 1 to 10 corresponding to the spin state shown in figure A.1.

Figure A.1.

Figure A.1. Notation for thorium nucleus spin states used in this section.

Standard image

The average square $\mathcal {H}_{\mathrm {Th}}^2$ of the magnetic field at the position of the thorium nucleus is given by

Equation (A.21)

The relaxation rates Γij = −Rijij for optical coherences ρij due to interaction with the random magnetic field created by the surrounding fluorine spins can be calculated in the same manner as for a single fluorine spin above. These rates (for allowed magnetic dipole transitions) are equal to

Equation (A.22)

Here we used the values of the thorium nuclear magnetic moment in the ground state μg = 0.45μN and in the isomer state μe = −0.076μN, respectively [50]. The decoherence rates Γ3,8 = Γ4,9 have been used in the evaluations of the nuclear clock performance in section 5.

Appendix B.: Validity of analytic expressions for the frequency fluctuation

Here we compare the numerically optimized results for δf obtained in section 5 with the analytic expressions obtained in section 4 for the limiting cases of short and long interrogation times. For short t, δft−3/2; therefore we express δf·t3/2 as a function of R and compare these results with the results of the analytical formulae (37) and (39); see figure B.1(a). For long t we express δf as a function of R and compare this result with formulae (44) and (47).

Figure B.1.

Figure B.1. (a) Dependence of δf·t3/2 for short interrogation time t on the excitation rate R obtained by the optimization of (31) (curve 1), using approximate formulae (39) (line 2) and (37) (line 3). (b) Dependence of δf for long t on R obtained by the optimization of (31) (curve 1), using approximate formulae (47) (line 2) and (44) (line 3).

Standard image

We conclude that the approximate analytic formulae (37), (39), (44) and (47) describe well the behavior of the optimized frequency error δf of the nuclear clock for different limiting cases of short and long interrogation times and for weak and strong laser fields. Taking into account the quite smooth dependence of δf on the interrogation time (see figure 6), we conclude that one can use expressions (37), (39), (44) and (47) as a lower-bound estimation for the frequency fluctuation δf of the fluorescence spectroscopy interrogation scheme.

Footnotes

  • The thorium first, second, third and fourth ionization energies are 6.3 eV [30], 11.5 eV, 20 eV and 29 eV, respectively. Therefore, the expected energy of the isomer transition (7.8 eV) is above the ionization energy of the neutral thorium atom.

  • Although this condition is not entirely fulfilled for the relaxation of fluorine spins in the random magnetic field $\skew5\vec {\mathcal {H}}$ created by neighboring spins, the quite good correspondence with experimental results indicates the applicability of this theory. For the thorium nucleus interacting with $\skew5\vec {\mathcal {H}}$ , this condition is fulfilled much better due to the much lower magnetic moment compared to fluorine.

Please wait… references are loading.