\floatsetup

[table]capposition=top

Ramsey interferometry of nuclear spins in diamond
using stimulated Raman adiabatic passage

Sean Lourette [email protected] Department of Physics, University of California, Berkeley, California 94720, USA DEVCOM Army Research Laboratory, Adelphi, Maryland 20783, USA    Andrey Jarmola Department of Physics, University of California, Berkeley, California 94720, USA DEVCOM Army Research Laboratory, Adelphi, Maryland 20783, USA    Jabir Chathanathil DEVCOM Army Research Laboratory, Adelphi, Maryland 20783, USA    Sebastián C. Carrasco DEVCOM Army Research Laboratory, Adelphi, Maryland 20783, USA    Dmitry Budker Department of Physics, University of California, Berkeley, California 94720, USA Helmholtz-Institut Mainz, 55099 Mainz, Germany GSI Helmholtzzentrum für Schwerionenforschung GmbH, 64291 Darmstadt, Germany Johannes Gutenberg-Universität Mainz, 55128 Mainz, Germany    Svetlana A. Malinovskaya Helmholtz-Institut Mainz, 55099 Mainz, Germany GSI Helmholtzzentrum für Schwerionenforschung GmbH, 64291 Darmstadt, Germany Johannes Gutenberg-Universität Mainz, 55128 Mainz, Germany Department of Physics, Stevens Institute of Technology, Hoboken, NJ 07030 Technische Universität Darmstadt, Institut für Angewandte Physik, D-64289 Darmstadt, Germany    A. Glen Birdwell DEVCOM Army Research Laboratory, Adelphi, Maryland 20783, USA    Tony G. Ivanov DEVCOM Army Research Laboratory, Adelphi, Maryland 20783, USA    Vladimir S. Malinovsky DEVCOM Army Research Laboratory, Adelphi, Maryland 20783, USA
(July 22, 2024)
Abstract

We report the first experimental demonstration of stimulated Raman adiabatic passage (STIRAP) in nuclear-spin transitions of 14N within nitrogen-vacancy (NV) color centers in diamond. It is shown that the STIRAP technique suppresses the occupation of the intermediate state, which is a crucial factor for improvements in quantum sensing technology. Building on that advantage, we develop and implement a generalized version of the Ramsey interferometric scheme, employing half-STIRAP pulses to perform the necessary quantum-state manipulation with high fidelity. The enhanced robustness of the STIRAP-based Ramsey scheme to variations in the pulse parameters is experimentally demonstrated, showing good agreement with theoretical predictions. Our results pave the way for improving the long-term stability of diamond-based sensors, such as gyroscopes and frequency standards.

I Introduction

Quantum sensing signifies a groundbreaking development in detection methodologies. Leveraging the principles of quantum mechanics, it enables the creation of remarkably precise and sensitive detectors [1]. A typical quantum sensor involves the use of quantum particles (such as photons, atoms, or ions), the manipulation of these particles’ states, and the precise measurement of changes in those states. The highest precision atomic clocks employ a basic two-level (TL) atomic system as their core element, with their superior performance being directly linked to the ability to accurately control atomic states [2].

The TL atomic system driven by an electromagnetic field has proven to be surprisingly rich in physical phenomena [3]. The quantum control of the TL wave function has been extensively studied and successfully implemented in various technological applications. The primary mechanism for controlling a TL system relies on the pulse area theorem, which is evidenced by Rabi oscillations of the state population as a function of the area of the applied pulses  [3, 4, 5, 6].

However, it is often necessary to control more than just two states of a primary sensor particle. Even adding a third level to the system introduces multiple additional effects that cannot be fathomed in a simple TL particle. Examples include electromagnetically induced transparency (EIT) [7, 8] and stimulated Raman adiabatic passage (STIRAP) [9, 10, 11, 12, 13, 14, 15]. The latter allows robust transfer of populations among the three states even when one of the states is “lossy”. STIRAP has numerous applications across various fields of science [16]. The basic STIRAP scheme has been extended to multi-level systems [17], and to “fractional STIRAP” [18], which allows the control of coherence, and more recently to “chirped STIRAP” [19, 20], which, in addition to controlling coherence, achieves state manipulation with high spectral resolution. Another extension is “two-way STIRAP” [21], which enables robust swap of populations between two states. Quantum control of entanglement applying STIRAP and fractional STIRAP has been considered in [22, 23] showing robustness of the adiabatic method for entangled states generation.

In this work, we utilize pulse-area and STIRAP control methods to manipulate the quantum states of the 14N nuclear spins (I=1𝐼1I=1italic_I = 1) within nitrogen-vacancy (NV) color centers in diamond. Since the 14N nuclear spin states form a three-level ΛΛ\Lambdaroman_Λ system, it is natural to consider STIRAP pulse sequences to control spin dynamics. We develop an adiabatic extension of advanced Ramsey interferometry based on the STIRAP protocol. We compare the performance of conventional and STIRAP-based Ramsey and demonstrate the advantages of the latter, particularly its robustness against variations in pulse parameters. The presented results offer practical advantages and will facilitate further advancements in quantum sensing science and technology.

This study, along with previous work by others [24], demonstrates the successful application of STIRAP techniques in the field of nuclear magnetic resonance (NMR). Notably, this work reverses the typical trend, as NMR techniques are often adapted for use in atomic and molecular physics [25]. Here, we showcase the successful transfer of a technique from atomic and molecular physics to the NMR domain, highlighting its versatility and potential for cross-disciplinary innovation.

Refer to caption
Figure 1: STIRAP in N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spins intrinsic to NV centers (a) NV ground-state energy levels. STIRAP is performed on the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spin triplet in the mS=0subscript𝑚𝑆0m_{S}=0italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 0 electron spin state using RF pulses with frequencies near the transition frequencies f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (5.089 MHz)times5.089MHz($5.089\text{\,}\mathrm{M}\mathrm{H}\mathrm{z}$)( start_ARG 5.089 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG ) and f2subscript𝑓2f_{2}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (4.797 MHz)times4.797MHz($4.797\text{\,}\mathrm{M}\mathrm{H}\mathrm{z}$)( start_ARG 4.797 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG ). The readout of the nuclear spin is performed either through ODMR using microwaves on the |mS=0|mS=+1ketsubscript𝑚𝑆0ketsubscript𝑚𝑆1\ket{m_{S}=0}\Leftrightarrow\ket{m_{S}=+1}| start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 0 end_ARG ⟩ ⇔ | start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = + 1 end_ARG ⟩ transitions or through direct optical readout. (b) Energy level diagram of the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear-spin sublevels within |mS=0ketsubscript𝑚𝑆0\ket{m_{S}=0}| start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 0 end_ARG ⟩ depicting the Rabi frequencies for the Stokes (Ωs(t))subscriptΩ𝑠𝑡\big{(}\Omega_{s}(t)\big{)}( roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) ) and pump (Ωp(t))subscriptΩ𝑝𝑡\big{(}\Omega_{p}(t)\big{)}( roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) ) pulses and the one-photon detuning ΔΔ\Deltaroman_Δ. (c) Pulse sequence diagram showing the readout of the nuclear spin state using ODMR. (d) Pulse shapes for pump and Stokes pulses for STIRAP and 2×STIRAP2STIRAP2\times\mathrm{STIRAP}2 × roman_STIRAP. (e) Fidelity of STIRAP, demonstrated for Δ=0Δ0\Delta=0roman_Δ = 0. Pulsed ODMR is used to measure the populations of the nuclear-spin sublevels after initialization, STIRAP, and 2×STIRAP2STIRAP2\times\mathrm{STIRAP}2 × roman_STIRAP. Markers – experimental measurements, solid lines – Lorentzian fit, vertical dashed lines – positions of the three hyperfine components.

II STIRAP in nuclear spins in diamond

To demonstrate the STIRAP population transfer between 14N nuclear spin states, we use a custom-built epifluorescence microscopy setup (described in [26]) to measure optically detected magnetic resonances (ODMRs) in an ensemble of NV centers. The diamond used for our experiments is chemical-vapor-deposition (CVD) grown, C12superscriptC12{}^{12}\mathrm{C}start_FLOATSUPERSCRIPT 12 end_FLOATSUPERSCRIPT roman_C enriched (99.99% C12superscriptC12{}^{12}\mathrm{C}start_FLOATSUPERSCRIPT 12 end_FLOATSUPERSCRIPT roman_C) [110] cut diamond plate, with initial nitrogen concentration of 13 ppmsimilar-toabsenttimes13ppm\sim$13\text{\,}\mathrm{p}\mathrm{p}\mathrm{m}$∼ start_ARG 13 end_ARG start_ARG times end_ARG start_ARG roman_ppm end_ARG and NV concentration of 4 ppmsimilar-toabsenttimes4ppm\sim$4\text{\,}\mathrm{p}\mathrm{p}\mathrm{m}$∼ start_ARG 4 end_ARG start_ARG times end_ARG start_ARG roman_ppm end_ARG. A bias magnetic field B𝐵Bitalic_B (480 Gtimes480G480\text{\,}\mathrm{G}start_ARG 480 end_ARG start_ARG times end_ARG start_ARG roman_G end_ARG) was applied along one of the NV axes using Halbach array composed of temperature-compensated samarium-cobalt (SmCo) magnets, designed to minimize the magnetic field gradients across the volume (\qtyproduct50x50x150\micromsimilar-toabsent\qtyproduct50𝑥50𝑥150\micro𝑚\sim\qtyproduct{50x50x150}{\micro m}∼ 50 italic_x 50 italic_x 150 italic_m) of interrogated NV centers (see Appendix A).

Figure 1(a) shows the electron NV ground state triplet energy levels (mS=0,±1subscript𝑚𝑆0plus-or-minus1m_{S}=0,\pm 1italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 0 , ± 1) and the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N hyperfine sublevels (mI=0,±1subscript𝑚𝐼0plus-or-minus1m_{I}=0,\pm 1italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 0 , ± 1) of |mS=0ketsubscript𝑚𝑆0\ket{m_{S}=0}| start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 0 end_ARG ⟩ and |mS=+1ketsubscript𝑚𝑆1\ket{m_{S}=+1}| start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = + 1 end_ARG ⟩ for a magnetic field B=480 G𝐵times480GB=$480\text{\,}\mathrm{G}$italic_B = start_ARG 480 end_ARG start_ARG times end_ARG start_ARG roman_G end_ARG aligned to the NV axis. At this magnetic field, the electron and nuclear spins are optically polarized into the |mS,mI=|0,+1ketsubscript𝑚𝑆subscript𝑚𝐼ket01\ket{m_{S},m_{I}}=\ket{0,+1}| start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ⟩ = | start_ARG 0 , + 1 end_ARG ⟩ state, as a result of the excited-state level anticrossing (ESLAC). This phenomenon is caused by an electro-nuclear spin-conserving flip-flop interaction, allowing the electron-spin polarization to be effectively transferred to the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spin, and is explained in [27, 28, 29, 30].

In this work, STIRAP (Fig. 1(a), orange box) is performed on the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spins that are intrinsic to the NV center. Population is transferred from mI=+1subscript𝑚𝐼1m_{I}=+1italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = + 1 to mI=1subscript𝑚𝐼1m_{I}=-1italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = - 1 in the mS=0subscript𝑚𝑆0m_{S}=0italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 0 manifold using two radio-frequency (RF) fields with frequencies ωpsubscript𝜔𝑝\omega_{p}italic_ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT (of about f1=subscript𝑓1absentf_{1}=italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 5.089 MHztimes5.089MHz5.089\text{\,}\mathrm{M}\mathrm{H}\mathrm{z}start_ARG 5.089 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG) and ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT (of about f2=subscript𝑓2absentf_{2}=italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 4.797 MHztimes4.797MHz4.797\text{\,}\mathrm{M}\mathrm{H}\mathrm{z}start_ARG 4.797 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG), and amplitudes Ωp(t)subscriptΩ𝑝𝑡\Omega_{p}(t)roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) and Ωs(t)subscriptΩ𝑠𝑡\Omega_{s}(t)roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ), respectively.

The population of the three nuclear-spin sublevels is determined using optically detected magnetic resonance (ODMR) techniques [31, 32]. There are three microwave (MW) transitions (Fig. 1(a), cyan box) between |mS=0ketsubscript𝑚𝑆0\ket{m_{S}=0}| start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 0 end_ARG ⟩ and |mS=+1ketsubscript𝑚𝑆1\ket{m_{S}=+1}| start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = + 1 end_ARG ⟩ that follow ΔmI=0Δsubscript𝑚𝐼0\Delta m_{I}=0roman_Δ italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 0 selection rules, each transition corresponding to a particular nuclear spin state (mIsubscript𝑚𝐼m_{I}italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT). The three transitions result in the appearance of three resonances (4.2 GHzsimilar-toabsenttimes4.2GHz\sim$4.2\text{\,}\mathrm{G}\mathrm{H}\mathrm{z}$∼ start_ARG 4.2 end_ARG start_ARG times end_ARG start_ARG roman_GHz end_ARG) in the ODMR spectrum that are used to determine the population of each mIsubscript𝑚𝐼m_{I}italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT state. The relative populations of the nuclear spin state can also be measured using direct optical readout, which is feasible near ESLAC for reasons similar to those responsible for optical polarization [33].

II.1 Sketch of STIRAP theory

The Hamiltonian of the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spin states in the field-interaction representation under the rotating wave approximation (RWA) in the case of two-photon resonance has the form

H(t)=2(0Ωp(t)0Ωp(t)2ΔΩs(t)0Ωs(t)0),𝐻𝑡Planck-constant-over-2-pi20subscriptΩ𝑝𝑡0subscriptΩ𝑝𝑡2ΔsubscriptΩ𝑠𝑡0subscriptΩ𝑠𝑡0\displaystyle H(t)=\frac{\hbar}{2}\left(\begin{array}[]{ccc}0&\Omega_{p}(t)&0% \\ \Omega_{p}(t)&-2\Delta&\Omega_{s}(t)\\ 0&\Omega_{s}(t)&0\end{array}\right)\,,italic_H ( italic_t ) = divide start_ARG roman_ℏ end_ARG start_ARG 2 end_ARG ( start_ARRAY start_ROW start_CELL 0 end_CELL start_CELL roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL - 2 roman_Δ end_CELL start_CELL roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL 0 end_CELL end_ROW end_ARRAY ) , (4)

where Δ=ωpω1=ωsω2Δsubscript𝜔𝑝subscript𝜔1subscript𝜔𝑠subscript𝜔2\Delta=\omega_{p}-\omega_{1}=\omega_{s}-\omega_{2}roman_Δ = italic_ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is the one-photon detuning shown in Fig. 1(b), ωp,ssubscript𝜔𝑝𝑠\omega_{p,s}italic_ω start_POSTSUBSCRIPT italic_p , italic_s end_POSTSUBSCRIPT are the frequencies of the pump and Stokes RF-fields, ω1,2=2πf1,2subscript𝜔122𝜋subscript𝑓12\omega_{1,2}=2\pi f_{1,2}italic_ω start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT = 2 italic_π italic_f start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT are the frequencies of |mI=1|mI=0ketsubscript𝑚𝐼1ketsubscript𝑚𝐼0\ket{m_{I}=1}\leftrightarrow\ket{m_{I}=0}| start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 1 end_ARG ⟩ ↔ | start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 0 end_ARG ⟩ and |mI=0|mI=1ketsubscript𝑚𝐼0ketsubscript𝑚𝐼1\ket{m_{I}=0}\leftrightarrow\ket{m_{I}=-1}| start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 0 end_ARG ⟩ ↔ | start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = - 1 end_ARG ⟩ transitions depicted in Fig. 1(a), Ωp,s(t)=γnBp,s(t)subscriptΩ𝑝𝑠𝑡subscript𝛾𝑛subscript𝐵𝑝𝑠𝑡\Omega_{p,s}(t)=\gamma_{n}B_{p,s}(t)roman_Ω start_POSTSUBSCRIPT italic_p , italic_s end_POSTSUBSCRIPT ( italic_t ) = italic_γ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_p , italic_s end_POSTSUBSCRIPT ( italic_t ) are the Rabi frequency envelopes of the pump and Stokes pulses. Here, Bp,s(t)subscript𝐵𝑝𝑠𝑡B_{p,s}(t)italic_B start_POSTSUBSCRIPT italic_p , italic_s end_POSTSUBSCRIPT ( italic_t ) refers to the amplitudes of the oscillations of the transverse magnetic field (perpendicular to the NV quantization axis) produced by the pump and Stokes fields, and γn/2π=307.59(3) Hz/Gsubscript𝛾𝑛2𝜋timesuncertain307.593HzG\gamma_{n}/2\pi=$307.59(3)\text{\,}\mathrm{H}\mathrm{z}\mathrm{/}\mathrm{G}$italic_γ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT / 2 italic_π = start_ARG start_ARG 307.59 end_ARG start_ARG ( 3 ) end_ARG end_ARG start_ARG times end_ARG start_ARG roman_Hz / roman_G end_ARG [34] is the gyromagnetic ratio of the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spin in diamond.

The commonly accepted mechanism of the STIRAP protocol by which population is transferred from the initial state |1ket1\ket{1}| start_ARG 1 end_ARG ⟩ to the target state |1ket1\ket{-1}| start_ARG - 1 end_ARG ⟩ can be demonstrated in the adiabatic basis. Diagonalizing the Hamiltonian in Eq. (4) reveals the three eigenstates

D⟩ ​ = ​ (cosθ0 -sinθ) ​​,  —B_-⟩ ​ = ​ (sinθsinξcosξcosθsinξ) ​​,  —B_+⟩ ​ = ​ (sinθcosξ-sinξcosθcosξ) ​​,

(5)

where time dependence of all terms has been omitted to simplify the notation, tanθ(t)=Ωp(t)/Ωs(t)𝜃𝑡subscriptΩ𝑝𝑡subscriptΩ𝑠𝑡\tan\theta(t)=\Omega_{p}(t)/\Omega_{s}(t)roman_tan italic_θ ( italic_t ) = roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) / roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) and tan2ξ(t)=Ωe(t)/Δ2𝜉𝑡subscriptΩ𝑒𝑡Δ\tan 2\xi(t)=\Omega_{e}(t)/\Deltaroman_tan 2 italic_ξ ( italic_t ) = roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_t ) / roman_Δ define the mixing angles θ(t)𝜃𝑡\theta(t)italic_θ ( italic_t ) and ξ(t)𝜉𝑡\xi(t)italic_ξ ( italic_t ), and Ωe(t)=Ωp2(t)+Ωs2(t)subscriptΩ𝑒𝑡superscriptsubscriptΩ𝑝2𝑡superscriptsubscriptΩ𝑠2𝑡\Omega_{e}(t)=\sqrt{\Omega_{p}^{2}(t)+\Omega_{s}^{2}(t)}roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_t ) = square-root start_ARG roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_t ) + roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_t ) end_ARG is the effective Rabi frequency [35]. We define pulse area as 𝒜=Ωe(t)dt𝒜subscriptΩ𝑒𝑡differential-d𝑡\mathcal{A}=\int\Omega_{e}(t)\,\mathrm{d}tcaligraphic_A = ∫ roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_t ) roman_d italic_t.

From the expression of the so-called dark state |𝒟(t)ket𝒟𝑡\ket{\mathcal{D}(t)}| start_ARG caligraphic_D ( italic_t ) end_ARG ⟩ we can see that the “counterintuitive” or SP pulse sequence (the Stokes pulse precedes the pump pulse and it is turned off before the pump pulse ends) provides a possibility to transfer population from the state |1ket1\ket{1}| start_ARG 1 end_ARG ⟩ to the state |1ket1\ket{-1}| start_ARG - 1 end_ARG ⟩. Indeed, if Ωp(t)/Ωs(t)0subscriptΩ𝑝𝑡subscriptΩ𝑠𝑡0\Omega_{p}(t)/\Omega_{s}(t)\rightarrow 0roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) / roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) → 0 (θ(t)0𝜃𝑡0\theta(t)\rightarrow 0italic_θ ( italic_t ) → 0) the eigenstate |𝒟(t)ket𝒟𝑡\ket{\mathcal{D}(t)}| start_ARG caligraphic_D ( italic_t ) end_ARG ⟩ in Eq. (5) correlates with state |1ket1\ket{1}| start_ARG 1 end_ARG ⟩, while if Ωs(t)/Ωp(t)0subscriptΩ𝑠𝑡subscriptΩ𝑝𝑡0\Omega_{s}(t)/\Omega_{p}(t)\rightarrow 0roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) / roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) → 0 (θ(t)π/2𝜃𝑡𝜋2\theta(t)\rightarrow\pi/2italic_θ ( italic_t ) → italic_π / 2) the eigenstate correlates with state |1ket1\ket{-1}| start_ARG - 1 end_ARG ⟩. To satisfy the transfer adiabaticity (to guarantee the system dynamics take place only in the zero eigenstate |𝒟(t)ket𝒟𝑡\ket{\mathcal{D}(t)}| start_ARG caligraphic_D ( italic_t ) end_ARG ⟩) and minimize residual population of the intermediate state |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ the pulses should be sufficiently strong and have substantial overlap [9].

II.2 STIRAP demonstration

Figure 1(c) illustrates the pulse sequence utilized for manipulating nuclear spin states and subsequently measuring the populations of these states. First, the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spins are optically initialized (i.e. polarized) into mI=+1subscript𝑚𝐼1m_{I}=+1italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = + 1 using a 532 nmtimes532nm532\text{\,}\mathrm{n}\mathrm{m}start_ARG 532 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG laser pulse, making use of the ESLAC at 480 Gtimes480G480\text{\,}\mathrm{G}start_ARG 480 end_ARG start_ARG times end_ARG start_ARG roman_G end_ARG [27, 28, 29, 30]. Next, the state of nuclear spin is manipulated (e.g. STIRAP, 2×STIRAP2STIRAP2\times\mathrm{STIRAP}2 × roman_STIRAP) using RF pulses with frequencies ωpsubscript𝜔𝑝\omega_{p}italic_ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. Then the population of each nuclear spin state is measured using pulsed ODMR [31], in which the change in fluorescence is measured with and without a mapping microwave π𝜋\piitalic_π-pulse (rectangular pulse of duration 2 µstimes2µs2\text{\,}\mathrm{\SIUnitSymbolMicro}\mathrm{s}start_ARG 2 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_s end_ARG) whose frequency is scanned through the |mS=0|mS=+1ketsubscript𝑚𝑆0ketsubscript𝑚𝑆1\ket{m_{S}=0}\Leftrightarrow\ket{m_{S}=+1}| start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 0 end_ARG ⟩ ⇔ | start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = + 1 end_ARG ⟩ transition.

To implement STIRAP transfer, we use the pump and Stokes pulses of the Blackman shape, defining the time dependent Rabi frequencies as Ωp,s(t)=Ω0wB(ttd/2)subscriptΩ𝑝𝑠𝑡subscriptΩ0subscript𝑤𝐵minus-or-plus𝑡subscript𝑡𝑑2\Omega_{p,s}(t)=\Omega_{0}w_{B}(t\mp t_{d}/2)roman_Ω start_POSTSUBSCRIPT italic_p , italic_s end_POSTSUBSCRIPT ( italic_t ) = roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_t ∓ italic_t start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / 2 ), where

wB(t)=sin2(πt/T)0.16sin2(2πt/T),subscript𝑤𝐵𝑡superscript2𝜋𝑡𝑇0.16superscript22𝜋𝑡𝑇\displaystyle w_{B}(t)=\sin^{2}\left(\pi t/T\right)-0.16\sin^{2}\left(2\pi t/T% \right)\,,italic_w start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_t ) = roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_π italic_t / italic_T ) - 0.16 roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 2 italic_π italic_t / italic_T ) , (6)

for 0tT0𝑡𝑇0\leq t\leq T0 ≤ italic_t ≤ italic_T, Ω0subscriptΩ0\Omega_{0}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the pulse amplitude, T𝑇Titalic_T is the individual pulse duration, and tdsubscript𝑡𝑑t_{d}italic_t start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT is the offset parameter controlling the time delay between pulses. For all measurements performed in this work, we use td=0.25T=0.2tpsubscript𝑡𝑑0.25𝑇0.2subscript𝑡𝑝t_{d}=0.25T=0.2t_{p}italic_t start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 0.25 italic_T = 0.2 italic_t start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, where tpsubscript𝑡𝑝t_{p}italic_t start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is defined to be the composite pulse duration, to minimize the nonadiabatic coupling between the dark state |𝒟(t)ket𝒟𝑡\ket{\mathcal{D}(t)}| start_ARG caligraphic_D ( italic_t ) end_ARG ⟩ and the bright states |±(t)ketsubscriptplus-or-minus𝑡\ket{\mathcal{B}_{\pm}(t)}| start_ARG caligraphic_B start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ( italic_t ) end_ARG ⟩ of the Hamiltonian in Eq. (4). For this particular construction of STIRAP pulses, the pulse area can be expressed as 𝒜0.5093Ωpeaktp𝒜0.5093subscriptΩ𝑝𝑒𝑎𝑘subscript𝑡𝑝\mathcal{A}\approx 0.5093\,\Omega_{peak}\,t_{p}caligraphic_A ≈ 0.5093 roman_Ω start_POSTSUBSCRIPT italic_p italic_e italic_a italic_k end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, where Ωpeak=|Ωe(t)|max1.094Ω0subscriptΩ𝑝𝑒𝑎𝑘subscriptsubscriptΩ𝑒𝑡max1.094subscriptΩ0\Omega_{peak}=\left|\Omega_{e}(t)\right|_{\mathrm{max}}\approx 1.094\,\Omega_{0}roman_Ω start_POSTSUBSCRIPT italic_p italic_e italic_a italic_k end_POSTSUBSCRIPT = | roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_t ) | start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ≈ 1.094 roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the peak effective Rabi frequency.

To demonstrate the fidelity of STIRAP, we measure the pulsed-ODMR spectra (Fig. 1(e), top to bottom): (i) after initialization (without RF pulses), in which the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spins are optically polarized into |mI=+1ketsubscript𝑚𝐼1\ket{m_{I}=+1}| start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = + 1 end_ARG ⟩, (ii) after STIRAP transfer (Fig. 1(d)-top), when the population is transferred from |mI=+1ketsubscript𝑚𝐼1\ket{m_{I}=+1}| start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = + 1 end_ARG ⟩ to |mI=1ketsubscript𝑚𝐼1\ket{m_{I}=-1}| start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = - 1 end_ARG ⟩, and (iii) after 2×STIRAP2STIRAP2\times\mathrm{STIRAP}2 × roman_STIRAP transfer (Fig. 1(d)-bottom), when the population is transferred to |mI=1ketsubscript𝑚𝐼1\ket{m_{I}=-1}| start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = - 1 end_ARG ⟩ (the first STIRAP) and then back to |mI=+1ketsubscript𝑚𝐼1\ket{m_{I}=+1}| start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = + 1 end_ARG ⟩ (second, or reverse STIRAP). Note that the ODMR signal after STIRAP appears to be weaker than the signal after initialization (compare the top and the middle frames in Fig. 1(e)). This, however, does not indicate the loss of the population, but rather it reflects the difference of the relative brightness of the nuclear states [33]. Indeed, after STIRAP is performed a second time the signal largely returns to its original size, 95% of its initially polarized value.

The dynamics of the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spin during STIRAP is measured by varying t/tp𝑡subscript𝑡𝑝t/t_{p}italic_t / italic_t start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, the fraction of the applied STIRAP pulse (Fig. 2(a)), and subsequently measuring populations of |mIketsubscript𝑚𝐼\ket{m_{I}}| start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ⟩ using pulsed ODMR. In contrast with measurements shown in Fig. 1(e), where the microwave frequency was scanned, here the microwave frequency was sequentially fixed to each of the three |mS=0,mI|mS=+1,mIketsubscript𝑚𝑆0subscript𝑚𝐼ketsubscript𝑚𝑆1subscript𝑚𝐼\ket{m_{S}=0,m_{I}}\Leftrightarrow\ket{m_{S}=+1,m_{I}}| start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 0 , italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ⟩ ⇔ | start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = + 1 , italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ⟩ transitions corresponding to the three hyperfine components. This provides sufficient information to uniquely determine the nuclear spin state population.

Refer to caption
Figure 2: STIRAP Dynamics. (a) Amplitudes of the applied truncated RF STIRAP Stokes and pump pulses, shown here for t/tp=0.6,𝑡subscript𝑡𝑝0.6t/t_{p}=0.6,italic_t / italic_t start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 0.6 , which are used to determine the time-evolution of the 14N nuclear-spin-state population. The pulse area for the full non-truncated sequence is 𝒜/25.6π𝒜25.6𝜋\mathcal{A}/2\approx 5.6\picaligraphic_A / 2 ≈ 5.6 italic_π. (b),(c) The time-evolution of the nuclear spin state during STIRAP for Δ/2π=\qtylist0;20kHzΔ2𝜋\qtylist020𝑘𝐻𝑧\Delta/2\pi=\qtylist{0;20}{kHz}roman_Δ / 2 italic_π = 0 ; 20 italic_k italic_H italic_z, respectively. (d) The same as (a), but with the ordering of Stokes and pump pulse swapped (“intuitive” or PS ordering). (e),(f) The time-evolution of the nuclear spin state during the PS pulse sequence for Δ/2π=\qtylist0;20kHzΔ2𝜋\qtylist020𝑘𝐻𝑧\Delta/2\pi=\qtylist{0;20}{kHz}roman_Δ / 2 italic_π = 0 ; 20 italic_k italic_H italic_z, respectively. Markers – experimental measurements, solid lines – theoretical model.
Refer to caption
Figure 3: Ramsey interferometry of 14N nuclear spins using STIRAP. (a) The pulse sequence for nuclear double-quantum (DQ) 4-Ramsey. (b) The pulse sequence for nuclear STIRAP 4-Ramsey, using two half-STIRAP pulses. For both (a) and (b): the duration of the RF pulses are drawn to scale, and the final pulse is phase cycled according to the inset. (c,d) The experimentally measured Ramsey fringes obtained by scanning τ𝜏\tauitalic_τ using the pulse sequences shown in (a) and (b) respectively. The four panels in (c) and (d) follow a similar structure: (Top-left) Ramsey fringes obtained by scanning τ𝜏\tauitalic_τ, for all four phase combinations and (Top-right) corresponding Fourier transforms. (Bottom-left) Linear combination of individual Ramsey measurements with different phases and (Bottom-right) corresponding Fourier transforms.

Figure 2(b),(c) shows the experimentally measured (markers) and theoretically modeled (lines) time evolution of the nuclear spin state population during STIRAP transfer for resonant (Δ=0Δ0\Delta=0roman_Δ = 0) and off-resonant (Δ/2π=20Δ2𝜋20\Delta/2\pi=20roman_Δ / 2 italic_π = 20 kHz) conditions. In both instances, the population is adiabatically transferred from |mI=+1ketsubscript𝑚𝐼1\ket{m_{I}=+1}| start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = + 1 end_ARG ⟩ to |mI=1ketsubscript𝑚𝐼1\ket{m_{I}=-1}| start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = - 1 end_ARG ⟩ through the dark state |𝒟(t)ket𝒟𝑡\ket{\mathcal{D}(t)}| start_ARG caligraphic_D ( italic_t ) end_ARG ⟩ (Eq. (5)), with a negligible transient population in the |mI=0ketsubscript𝑚𝐼0\ket{m_{I}=0}| start_ARG italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 0 end_ARG ⟩ state.

The same measurements are repeated with the PS ordering of the pulses (pump then Stokes, as depicted in Fig. 2(d)) resulting in nuclear spin state populations shown in Fig. 2(e) and Fig. 2(f) for Δ=0Δ0\Delta=0roman_Δ = 0 and Δ/2π=20Δ2𝜋20\Delta/2\pi=20roman_Δ / 2 italic_π = 20 kHz respectively. For Δ/2π=20 kHzΔ2𝜋times20kHz\Delta/2\pi=$20\text{\,}\mathrm{k}\mathrm{H}\mathrm{z}$roman_Δ / 2 italic_π = start_ARG 20 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG, there is near-complete population transfer to |1ket1\ket{-1}| start_ARG - 1 end_ARG ⟩, while for Δ=0Δ0\Delta=0roman_Δ = 0, the population is transferred to a nearly equal superposition of |1ket1\ket{-1}| start_ARG - 1 end_ARG ⟩ and |0ket0\ket{0}| start_ARG 0 end_ARG ⟩. These results can be understood in the adiabatic basis (in other words, by performing the dressed state analysis): in the case of the SP ordering, population is transferred exclusively through the dark state |𝒟(t)ket𝒟𝑡\ket{\mathcal{D}(t)}| start_ARG caligraphic_D ( italic_t ) end_ARG ⟩, but for the PS ordering the population is instead transferred through the two bright states |±(t)ketsubscriptplus-or-minus𝑡\ket{\mathcal{B}_{\pm}(t)}| start_ARG caligraphic_B start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ( italic_t ) end_ARG ⟩. For the PS ordering, the populations of these dressed states |(t)ketsubscript𝑡\ket{\mathcal{B}_{-}(t)}| start_ARG caligraphic_B start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ( italic_t ) end_ARG ⟩ and |+(t)ketsubscript𝑡\ket{\mathcal{B}_{+}(t)}| start_ARG caligraphic_B start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( italic_t ) end_ARG ⟩ at the initial time are defined by sinξ(t=0)𝜉𝑡0\sin\xi(t=0)roman_sin italic_ξ ( italic_t = 0 ) and cosξ(t=0)𝜉𝑡0\cos\xi(t=0)roman_cos italic_ξ ( italic_t = 0 ) respectively. A deeper analysis shows that regardless of the value of ΔΔ\Deltaroman_Δ, the population of the state |1ket1\ket{1}| start_ARG 1 end_ARG ⟩ is always zero after the pulse is complete, assuming the adiabaticity condition is satisfied. The details of the population dynamics at intermediate times depend on the effective pulse area 𝒜𝒜\mathcal{A}caligraphic_A and the detuning value ΔΔ\Deltaroman_Δ.

The dephasing rate in our sample is insufficient to produce the observed decay of the oscillations shown in Fig. 2(f) on the timescale of the applied pulses. In fact, this decay is explained by a gradient in RF amplitude, and therefore in the effective Rabi frequency Ωe(t)subscriptΩ𝑒𝑡\Omega_{e}(t)roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_t ), across the interrogated volume of NV centers. We measured the distribution of Rabi frequencies by performing Fourier analysis on Rabi oscillations of the |±1|0ketplus-or-minus1ket0\ket{\pm 1}\leftrightarrow\ket{0}| start_ARG ± 1 end_ARG ⟩ ↔ | start_ARG 0 end_ARG ⟩ transitions (see Fig. 7 in the Appendix). We found the measured Rabi frequency distribution to be well described by a normal distribution. All theoretical results have been obtained by averaging over the Rabi frequency distribution (see Appendix D).

The modeling results in Fig. 2 are obtained using the pulse duration of tp=300 μssubscript𝑡𝑝times300𝜇st_{p}=$300\text{\,}\mu\mathrm{s}$italic_t start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = start_ARG 300 end_ARG start_ARG times end_ARG start_ARG italic_μ roman_s end_ARG (T=240μ𝑇240𝜇T=240\muitalic_T = 240 italic_μs, td=60μsubscript𝑡𝑑60𝜇t_{d}=60\muitalic_t start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 60 italic_μs), and the peak amplitude of the effective Rabi frequency ΩesubscriptΩ𝑒\Omega_{e}roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT equal to Ωe/2π=36.5 kHzdelimited-⟨⟩subscriptΩ𝑒2𝜋times36.5kHz\langle\Omega_{e}\rangle/2\pi=$36.5\text{\,}\mathrm{k}\mathrm{H}\mathrm{z}$⟨ roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ⟩ / 2 italic_π = start_ARG 36.5 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG and full width at half maximum of 0.164Ωe0.164delimited-⟨⟩subscriptΩ𝑒0.164\,\left<\Omega_{e}\right>0.164 ⟨ roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ⟩, which agree with the experimentally measured values of 36.1 kHztimes36.1kHz36.1\text{\,}\mathrm{k}\mathrm{H}\mathrm{z}start_ARG 36.1 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG and 0.166Ωe0.166delimited-⟨⟩subscriptΩ𝑒0.166~{}\langle\Omega_{e}\rangle0.166 ⟨ roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ⟩. We note the high frequency oscillations with a very small amplitude, visible in Fig. 2 on the modeling lines. These residual oscillations are a result of the ac Zeeman shifts from the pump field acting off-resonantly on |0|1ket0ket1\ket{0}\leftrightarrow\ket{-1}| start_ARG 0 end_ARG ⟩ ↔ | start_ARG - 1 end_ARG ⟩ and the Stokes field on |1|0ket1ket0\ket{1}\leftrightarrow\ket{0}| start_ARG 1 end_ARG ⟩ ↔ | start_ARG 0 end_ARG ⟩ (see Appendix D for the explanation).

III Ramsey interferometry using half-STIRAP

Ramsey interferometry is a technique commonly used in sensing applications to precisely measure the transition frequencies of a quantum system. In a basic Ramsey interferometry scheme, a pair of π/2𝜋2\pi/2italic_π / 2 pulses delayed by a free evolution interval is applied to an ensemble of identical TL systems. Ideally, the durations of the pulses are negligible, and the area of each pulse is exactly equal to π/2𝜋2\pi/2italic_π / 2. Under these conditions, the measured Ramsey fringes are not sensitive to the pulse parameters (such as detuning), allowing for the detection of an external perturbation to the transition frequency through the phase shift of the fringes. In reality, the pulses have finite durations and may have non-ideal areas. Additionally, a homogeneous excitation of the ensemble may pose an experimental challenge. As a result of the imperfections, the measured phase shifts are sensitive to the pulse parameters and must be taken into account during sensor development.

In rotation sensing with N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spins that are intrinsic to NV centers [26, 36], the double-quantum (DQ) transition frequency (|mS,mIketsubscript𝑚𝑆subscript𝑚𝐼\ket{m_{S},m_{I}}| start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ⟩: |0,+1|0,1ket01ket01\ket{0,+1}\Leftrightarrow\ket{0,-1}| start_ARG 0 , + 1 end_ARG ⟩ ⇔ | start_ARG 0 , - 1 end_ARG ⟩) is measured using Ramsey interferometry, in which a pair of rectangular π/2𝜋2\pi/2italic_π / 2 RF pulses are used (Fig. 3a). However, when using ensembles of NV centers, there is a distribution of Rabi frequencies arising from RF power gradients across the sensing volume, which limits the fidelity and robustness of the technique. To overcome these shortcomings, we develop a new Ramsey technique replacing the rectangular pulses with adiabatic pulses, (half-STIRAP, Fig. 3b). In the previous section, we demonstrate the STIRAP protocol showing the adiabatic manipulation of the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spin state population. Here, we address the phase sensitivity of adiabatic control by creating an adiabatic version of the DQ Ramsey scheme and compare the performance of both techniques.

The STIRAP process demonstrated in the previous section can be modified to create a superposition of the initial |1ket1\ket{1}| start_ARG 1 end_ARG ⟩ and the target state |1ket1\ket{-1}| start_ARG - 1 end_ARG ⟩. This is straightforward to see using the dark state |𝒟(t)ket𝒟𝑡\ket{\mathcal{D}(t)}| start_ARG caligraphic_D ( italic_t ) end_ARG ⟩ defined in Eq. (5). When the turn-off condition is modified such that the pump and Stokes pulses satisfy Ωs(t)/Ωp(t)1subscriptΩ𝑠𝑡subscriptΩ𝑝𝑡1\Omega_{s}(t)/\Omega_{p}(t)\rightarrow 1roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) / roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) → 1, the superposition state (|1|1)/2ket1ket12(\ket{1}-\ket{-1})/\sqrt{2}( | start_ARG 1 end_ARG ⟩ - | start_ARG - 1 end_ARG ⟩ ) / square-root start_ARG 2 end_ARG is created. This is achieved without changing Ωe(t)subscriptΩ𝑒𝑡\Omega_{e}(t)roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_t ) by simply using θ(t)/2𝜃𝑡2\theta(t)/2italic_θ ( italic_t ) / 2 (which is swept from 00 to π/4𝜋4\pi/4italic_π / 4 instead of 00 to π/2𝜋2\pi/2italic_π / 2) as the mixing angle. This results in new pulse shapes for Ωp(t)subscriptΩ𝑝𝑡\Omega_{p}(t)roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) and Ωs(t)subscriptΩ𝑠𝑡\Omega_{s}(t)roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) according to the following equations

Ωp(t)=Ωe(t)sin(12θ(t)),Ωs(t)=Ωe(t)cos(12θ(t)).formulae-sequencesubscriptΩ𝑝𝑡subscriptΩ𝑒𝑡12𝜃𝑡subscriptΩ𝑠𝑡subscriptΩ𝑒𝑡12𝜃𝑡\displaystyle\begin{split}\Omega_{p}(t)&=\Omega_{e}(t)\sin\left(\tfrac{1}{2}% \theta(t)\right)\,,\\ \Omega_{s}(t)&=\Omega_{e}(t)\cos\left(\tfrac{1}{2}\theta(t)\right)\,.\end{split}start_ROW start_CELL roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL = roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_t ) roman_sin ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_θ ( italic_t ) ) , end_CELL end_ROW start_ROW start_CELL roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL = roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_t ) roman_cos ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_θ ( italic_t ) ) . end_CELL end_ROW (7)

We call this protocol half-STIRAP (H-STIRAP). It is interesting to note that the non-adiabatic coupling θ˙˙𝜃\dot{\theta}over˙ start_ARG italic_θ end_ARG in the H-STIRAP protocol is exactly half of that in the non-adiabatic coupling in the STIRAP protocol [10]. When using H-STIRAP pulses to perform Ramsey interferometry, the second H-STIRAP pulse is time reversed (see Appendix B).

III.1 Time dependence of Ramsey signal

Figure 3(a) shows the DQ Ramsey interferometry pulse sequence, which is described as follows. The N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spins are polarized into the |+1ket1\ket{+1}| start_ARG + 1 end_ARG ⟩ state using a green laser pulse. Next, the spin state is prepared in the superposition |ψ=(|+1+eiϕ|1)/2ket𝜓ket1superscript𝑒𝑖italic-ϕket12\ket{\psi}=\left(\ket{+1}+e^{i\phi}\ket{-1}\right)/\sqrt{2}| start_ARG italic_ψ end_ARG ⟩ = ( | start_ARG + 1 end_ARG ⟩ + italic_e start_POSTSUPERSCRIPT italic_i italic_ϕ end_POSTSUPERSCRIPT | start_ARG - 1 end_ARG ⟩ ) / square-root start_ARG 2 end_ARG using two RF pulses: a π𝜋\piitalic_π-pulse on f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (duration π/Ωp18 µs𝜋subscriptΩ𝑝times18µs\pi/\Omega_{p}\approx$18\text{\,}\mathrm{\SIUnitSymbolMicro}\mathrm{s}$italic_π / roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ≈ start_ARG 18 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_s end_ARG, frequency ωpsubscript𝜔𝑝\omega_{p}italic_ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT), which transfers population from |+1ket1\ket{+1}| start_ARG + 1 end_ARG ⟩ to |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ and a two-tone pulse (duration π/Ωe13 µs𝜋subscriptΩ𝑒times13µs\pi/\Omega_{e}\approx$13\text{\,}\mathrm{\SIUnitSymbolMicro}\mathrm{s}$italic_π / roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ≈ start_ARG 13 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_s end_ARG) with frequencies ωpsubscript𝜔𝑝\omega_{p}italic_ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, which transfers the state from |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ to |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩. After a free precession time interval τ𝜏\tauitalic_τ, a second (identical) two-tone RF pulse is then used to convert the accumulated phase into a population difference, which is read out optically. Figure 3(b) shows the adiabatic version of the DQ Ramsey interferometry pulse sequence, using half-STIRAP RF pulses whose construction is defined in Eq. (7). The first half-STRAP pulse adiabatically transfers the spin state from |+1ket1\ket{+1}| start_ARG + 1 end_ARG ⟩ to |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩, and a reverse half-STIRAP pulse is used to project the relative phase into a population difference. For both pulse sequences (Fig. 3(a) and Fig. 3(b)), we employ a four-phase measurement (4-Ramsey) [26, 31], in which the phases of the final pulse (ωpsubscript𝜔𝑝\omega_{p}italic_ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT) are inverted (+xxmaps-to𝑥𝑥+x\mapsto-x+ italic_x ↦ - italic_x), cycling through all four possible combinations (Fig. 3(a) inset: R1subscript𝑅1R_{1}italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, R2subscript𝑅2R_{2}italic_R start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, R3subscript𝑅3R_{3}italic_R start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, R4subscript𝑅4R_{4}italic_R start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT and Fig. 3(b) inset: S1subscript𝑆1S_{1}italic_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, S2subscript𝑆2S_{2}italic_S start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, S3subscript𝑆3S_{3}italic_S start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, S4subscript𝑆4S_{4}italic_S start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT).

Figure 3c (and Fig. 3d) shows optically detected Ramsey fringes from the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spin DQ transition, obtained by scanning τ𝜏\tauitalic_τ using the pulse sequence shown in Fig. 3a (and Fig. 3b). This scanning of τ𝜏\tauitalic_τ is performed in the lab frame, rather than in the rotating frame, resulting in oscillations that are detected at the transition frequency f1f2subscript𝑓1subscript𝑓2f_{1}-f_{2}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. In fact, scanning τ𝜏\tauitalic_τ in the rotating frame would produce no oscillations, since we have no two-photon detuning (ΔpΔs=0subscriptΔ𝑝subscriptΔ𝑠0\Delta_{p}-\Delta_{s}=0roman_Δ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - roman_Δ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 0). The top-left panels of both Fig. 3c and Fig. 3d show the Ramsey fringes for each of the four phase combinations for DQ 4-Ramsey (R1subscript𝑅1R_{1}italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, R2subscript𝑅2R_{2}italic_R start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, R3subscript𝑅3R_{3}italic_R start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, R4subscript𝑅4R_{4}italic_R start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) and STIRAP 4-Ramsey (S1subscript𝑆1S_{1}italic_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, S2subscript𝑆2S_{2}italic_S start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, S3subscript𝑆3S_{3}italic_S start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, S4subscript𝑆4S_{4}italic_S start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT), respectively. We find that when compared with individual DQ 4-Ramsey, the individual STIRAP 4-Ramsey measurements have a larger amplitude (30%similar-toabsentpercent30\sim 30\%∼ 30 %), reduced “ripple” (1/20similar-toabsent120\sim 1/20∼ 1 / 20), and an inverted phase. The Fourier transforms of these individual signals (top-right panel of Fig. 3c and Fig. 3d) reveal that the “ripple” occurs at frequencies f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and f2subscript𝑓2f_{2}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, corresponding to residual population in |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ due to pulse imperfections. Combining the four individual Ramsey measurements, (R=R1R2+R3R4𝑅subscript𝑅1subscript𝑅2subscript𝑅3subscript𝑅4R=R_{1}-R_{2}+R_{3}-R_{4}italic_R = italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) suppresses these residual signals, (bottom-left panel of Fig. 3c and Fig. 3d). The suppression of the residual signals is also observed in the Fourier-transform plots (bottom-right of Fig. 3c and Fig. 3d). The difference in amplitude (including the sign change) between DQ 4-Ramsey and STIRAP 4-Ramsey can be explained by the particular pair of states between which oscillations occur. In DQ 4-Ramsey, the amplitude is determined by the difference in brightness between |0ket0\ket{0}| start_ARG 0 end_ARG ⟩, and (|+1+eiϕ|1)/2ket1superscript𝑒𝑖italic-ϕket12\left(\ket{+1}+e^{i\phi}\ket{-1}\right)/\sqrt{2}( | start_ARG + 1 end_ARG ⟩ + italic_e start_POSTSUPERSCRIPT italic_i italic_ϕ end_POSTSUPERSCRIPT | start_ARG - 1 end_ARG ⟩ ) / square-root start_ARG 2 end_ARG, while in STIRAP 4-Ramsey, the amplitude is determined by the difference in brightness between |+1ket1\ket{+1}| start_ARG + 1 end_ARG ⟩, and (|0+eiϕ|1)/2ket0superscript𝑒𝑖italic-ϕket12\left(\ket{0}+e^{i\phi}\ket{-1}\right)/\sqrt{2}( | start_ARG 0 end_ARG ⟩ + italic_e start_POSTSUPERSCRIPT italic_i italic_ϕ end_POSTSUPERSCRIPT | start_ARG - 1 end_ARG ⟩ ) / square-root start_ARG 2 end_ARG.

Refer to caption
Figure 4: Robustness of DQ 4-Ramsey and STIRAP 4-Ramsey. The accumulated phase (ΦΦ\Phiroman_Φ) of DQ 4-Ramsey and STIRAP 4-Ramsey interferometer techniques is plotted as a function of two parameters: detuning (ΔΔ\Deltaroman_Δ) and effective Rabi frequency (ΩesubscriptΩ𝑒\Omega_{e}roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT). The left column shows experimentally measured results, and the right column shows theoretical predictions. For both techniques, the duration tpsubscript𝑡𝑝t_{p}italic_t start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and separation τ𝜏\tauitalic_τ of the pulses were fixed. For DQ 4-Ramsey (top row) τ=1.2 ms𝜏times1.2ms\tau=$1.2\text{\,}\mathrm{m}\mathrm{s}$italic_τ = start_ARG 1.2 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG, tp=13.9 µssubscript𝑡𝑝times13.9µst_{p}=$13.9\text{\,}\mathrm{\SIUnitSymbolMicro}\mathrm{s}$italic_t start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = start_ARG 13.9 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_s end_ARG, and the horizontal dashed line corresponds to tuned pulse areas for all three pulses (Fig. 3) 𝒜/2=π/2𝒜2𝜋2\mathcal{A}/2=\pi/2caligraphic_A / 2 = italic_π / 2, Ωe=2π×36.1 kHzsubscriptΩ𝑒2𝜋times36.1kHz\Omega_{e}=2\pi\times$36.1\text{\,}\mathrm{k}\mathrm{H}\mathrm{z}$roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 2 italic_π × start_ARG 36.1 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG. For STIRAP 4-Ramsey (bottom row) τ=0.8 ms𝜏times0.8ms\tau=$0.8\text{\,}\mathrm{m}\mathrm{s}$italic_τ = start_ARG 0.8 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG, tp=500 µssubscript𝑡𝑝times500µst_{p}=$500\text{\,}\mathrm{\SIUnitSymbolMicro}\mathrm{s}$italic_t start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = start_ARG 500 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_s end_ARG, and the horizontal dashed lines indicate where the pulse area is 𝒜/2=5π𝒜25𝜋\mathcal{A}/2=5\picaligraphic_A / 2 = 5 italic_π and 10π10𝜋10\pi10 italic_π.

III.2 Robustness

We characterize the robustness of the measurement of the DQ transition frequency (f1f2subscript𝑓1subscript𝑓2f_{1}-f_{2}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) against detuning (ΔΔ\Deltaroman_Δ) and Rabi frequency (ΩesubscriptΩ𝑒\Omega_{e}roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT) for the two techniques shown in Fig. 3: DQ 4-Ramsey and STIRAP 4-Ramsey. These two parameters are of particular interest for robustness because they are known to fluctuate when subjected to drifts in ambient temperature.

The accumulated phase ΦΦ\Phiroman_Φ was obtained by varying the phase of the final Ramsey pulse. More precisely, Φ=tan1(Q/I)Φsuperscript1𝑄𝐼\Phi=\tan^{-1}\big{(}Q/I\big{)}roman_Φ = roman_tan start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_Q / italic_I ), where I𝐼Iitalic_I is the measurement obtained as previously described in Fig. 3a and Fig. 3b, and Q𝑄Qitalic_Q is obtained in the same way, but shifting the DQ phase (relative phase between pump and Stokes) of the final RF pulse by 90 degrees (see Appendix C).

Figure 4 shows 2D color plots of the accumulated phase ΦΦ\Phiroman_Φ as a function of ΔΔ\Deltaroman_Δ and ΩesubscriptΩ𝑒\Omega_{e}roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, obtained using both theory and experiment for both DQ 4-Ramsey and STIRAP 4-Ramsey. It should be noted that the pulse durations, not pulse areas, are fixed for these plots. For DQ 4-Ramsey, the pulse durations were chosen to have optimal pulse areas (e.g. 𝒜/2=π/2𝒜2𝜋2\mathcal{A}/2=\pi/2caligraphic_A / 2 = italic_π / 2) when Ωe/2π=36.1 kHzsubscriptΩ𝑒2𝜋times36.1kHz\Omega_{e}/2\pi=$36.1\text{\,}\mathrm{k}\mathrm{H}\mathrm{z}$roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT / 2 italic_π = start_ARG 36.1 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG, indicated by the horizontal dashed line. For STIRAP 4-Ramsey, the accumulated phase ΦΦ\Phiroman_Φ is plotted against ΩpeaksubscriptΩ𝑝𝑒𝑎𝑘\Omega_{peak}roman_Ω start_POSTSUBSCRIPT italic_p italic_e italic_a italic_k end_POSTSUBSCRIPT, the peak value of Ωe(t)subscriptΩ𝑒𝑡\Omega_{e}(t)roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_t ).

As expected, we found that the range of conditions over which the interferometer performs well is broader for STIRAP 4-Ramsey than for DQ 4-Ramsey. It is interesting to note that the distribution of Rabi frequencies (due to RF gradients) actually improves the robustness of STIRAP 4-Ramsey, smoothing out the diagonal stripes, especially for large pulse areas.

We found that for DQ 4-Ramsey, the accumulated phase also depends greatly on the phase of the initial f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT π𝜋\piitalic_π-pulse. Because the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spins are initially polarized in the |+1ket1\ket{+1}| start_ARG + 1 end_ARG ⟩ state, STIRAP 4-Ramsey does not require any initial pulse, and benefits in robustness as a result. Note that in the electron-spin triplet of the NV center, the situation is reversed: polarization occurs in |mS=0ketsubscript𝑚𝑆0\ket{m_{S}=0}| start_ARG italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 0 end_ARG ⟩ and therefore STIRAP 4-Ramsey requires an initial π𝜋\piitalic_π pulse and DQ 4-Ramsey does not [12, 13].

Under our experimental conditions (Ωe/2π=36.1 kHzsubscriptΩ𝑒2𝜋times36.1kHz\Omega_{e}/2\pi=$36.1\text{\,}\mathrm{k}\mathrm{H}\mathrm{z}$roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT / 2 italic_π = start_ARG 36.1 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG, Δ=0Δ0\Delta=0roman_Δ = 0), we measure the dependence of the accumulated phase on detuning (first derivative) to be similar-to\sim 2.1 deg/kHztimes2.1degkHz2.1\text{\,}\mathrm{d}\mathrm{e}\mathrm{g}\mathrm{/}\mathrm{k}\mathrm{H}% \mathrm{z}start_ARG 2.1 end_ARG start_ARG times end_ARG start_ARG roman_deg / roman_kHz end_ARG for DQ 4-Ramsey. This corresponds to phase drift of similar-to\sim 1.3 mrad/Ktimes1.3mradK1.3\text{\,}\mathrm{m}\mathrm{rad}\mathrm{/}\mathrm{K}start_ARG 1.3 end_ARG start_ARG times end_ARG start_ARG roman_mrad / roman_K end_ARG, which follows from the temperature dependence of the f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and f2subscript𝑓2f_{2}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT transitions (35 Hz/Ktimes-35HzK-35\text{\,}\mathrm{H}\mathrm{z}\mathrm{/}\mathrm{K}start_ARG - 35 end_ARG start_ARG times end_ARG start_ARG roman_Hz / roman_K end_ARG[34]). As a result, we expect this level of phase drift to ultimately limit the bias stability of rotation sensing with N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spins. STIRAP 4-Ramsey was measured to be more robust than DQ4R against changes in ΔΔ\Deltaroman_Δ under the same experimental conditions by a factor of 5, which can be further improved by choosing a value of pulse area for which this dependence vanishes.

IV Conclusions and outlook

In this work, we have experimentally demonstrated STIRAP in the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear hyperfine manifold in the ground state of NV centers in diamond driven by RF pulses. A new method of measuring nuclear-spin-state population dynamics (specific for NV centers in diamond) was developed and implemented. We demonstrate a substantial suppression of the intermediate state population during STIRAP transfer between the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear spin states. The experimental results are in good agreement with the developed theory, taking into account both gradients of the RF field across the measured volume as well as ac Zeeman shifts related to off-resonant components of the RF field excitation.

Utilizing the advantages of the STIRAP protocol, we developed and implemented an advanced Ramsey interferometric method based on the half-STIRAP pulse sequence. Effectively, we replace the traditionally used π/2𝜋2\pi/2italic_π / 2-pulses in basic Ramsey schemes with half-STIRAP pulses. This modification allows for robust preparation of the state superposition and readout of the generated coherence. We compared the performance of STIRAP Ramsey with the performance of the DQ Ramsey scheme, which is based on the effective π/2𝜋2\pi/2italic_π / 2-pulse sequence, as in the standard Ramsey protocol. Our findings indicate that STIRAP Ramsey offers enhanced tolerance to moderate changes in applied pulse parameters, resulting in improved robustness of Ramsey signal phase and amplitude (contrast).

The results show that STIRAP and its variations can effectively manipulate nuclear state populations and coherences, driving progress in quantum sensing applications like rotation sensing [26, 36, 37], secondary frequency standards [38], quantum memory devices [39], and other quantum technologies. The technique’s robustness against experimental imperfections will benefit future sensing and spectroscopy advances. Additionally, the π/2𝜋2\pi/2italic_π / 2-pulse to half-STIRAP pulse conversion can be applied to external degrees of freedom in atom interferometry [40], inheriting the robustness benefits. This STIRAP Ramsey scheme can be extended to other spin and atomic systems, providing improvements in robustness.

Acknowledgments

The authors are grateful to Alexander Pines, Malcolm Levitt, Dieter Suter, Victor M. Acosta, and Genko Genov for stimulating discussions. S.L., A.J. S.C., and J.C. acknowledge support from the DEVCOM Army Research Laboratory under Cooperative Agreement No. W911NF-24-2-0050, No. W911NF-18-2-0037, No. W911NF-24-2-0044, and No. W911NF-22-2-0097. S.A.M. acknowledges support from the Office of Naval Research under Awards No. N00014-20-1-2086 and N00014-22-1-2374, as well as from the Helmholtz Institute Mainz Visitor Program and the Alexander von Humboldt Foundation. D.B. acknowledges support from the DEVCOM Army Research Laboratory under Cooperative Agreements No. W911NF2120180, W911NF2320093.

Appendix A Experimental setup details

A 0.79–numerical aperture aspheric condenser lens (Thorlabs, ACL25416U-A) was used to illuminate a spot of diameter 50 µmtimessimilar-toabsent50µm\sim 50\text{\,}\mathrm{\SIUnitSymbolMicro}\mathrm{m}start_ARG ∼ 50 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG on the diamond with 30 mWtimessimilar-toabsent30mW\sim 30\text{\,}\mathrm{m}\mathrm{W}start_ARG ∼ 30 end_ARG start_ARG times end_ARG start_ARG roman_mW end_ARG of 532 nm laser light (Coherent Verdi G5) and collect fluorescence. Laser pulses were generated passing a continuous wave beam through an acousto-optic modulator. The NV sensing volume is 0.4nLsimilar-toabsent0.4nL\sim 0.4\,$\mathrm{n}\mathrm{L}$∼ 0.4 roman_nL (\qtyproduct50x50x150\micromsimilar-toabsent\qtyproduct50𝑥50𝑥150\micro𝑚\sim\qtyproduct{50x50x150}{\micro m}∼ 50 italic_x 50 italic_x 150 italic_m), defined by the area of the incident laser beam and the length of its path through the diamond. The fluorescence was separated from the excitation light by a dichroic mirror, passed through a band-pass filter (650 to 800 nm), and detected by a free-space Si photodiode.

Radio-frequency (RF) pulses were generated by an RF synthesizer (Keysight 33512B) using arbitrary waveform generation and amplified using a broadband RF amplifier (Minicircuits LZY-22+). Microwave (MW) signals were generated by a MW synthesizer (Rohde & Schwarz SMW200A), formed into pulses using an RF switch (Minicircuits ZASW-2-50DR+), and amplified using a broadband MW amplifier (Minicircuits ZHL-50W-63+). RF and MW signals were combined using a diplexer (MarkiMicrowave DPX-0R5) and delivered using a 160-µmµm\mathrm{\SIUnitSymbolMicro}\mathrm{m}roman_µ roman_m-diameter copper wire placed on the diamond surface next to the optical focus.

A TTL pulse card (SpinCore, PBESR-PRO-500) was used to generate and synchronize the pulse sequence. A data acquisition card (National Instruments, USB-6361) was used to digitize experimentally measured signals.

Appendix B Half-STIRAP pulse construction

The construction of the half-STIRAP pulses is described in Eq. (7). Figure 5 shows the waveforms of the half-STIRAP pulses captured using an oscilloscope (green), in addition to plots of the Rabi frequencies of the pump (orange) and Stokes (blue) components.

Refer to caption
Figure 5: Waveforms of the half-STIRAP pulses used in STIRAP Ramsey. The waveforms of the applied RF pulses for STIRAP Ramsey as used in the experiment, shown here with a reduced value of pulse separation τ𝜏\tauitalic_τ, and for tp=200 µssubscript𝑡𝑝times200µst_{p}=$200\text{\,}\mathrm{\SIUnitSymbolMicro}\mathrm{s}$italic_t start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = start_ARG 200 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_s end_ARG. The waveforms of the half-STIRAP pulses are shown in green, captured using an oscilloscope. The visible oscillations are a result of the the pump and Stokes pulses beating against one another at their difference frequency. Plots of the pulse amplitudes (Ωp(t)subscriptΩ𝑝𝑡\Omega_{p}(t)roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) and Ωs(t)subscriptΩ𝑠𝑡\Omega_{s}(t)roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t )) of the pump (orange) and Stokes (blue) components are shown above. The orange and blue dashed lines show the pump and Stokes pulses for the original Blackman-shaped STIRAP pulses, respectively. Both sets of pulses share the same effective Rabi frequency Ωe(t)subscriptΩ𝑒𝑡\Omega_{e}(t)roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_t ), shown in black.
Refer to caption
Figure 6: Robustness of the amplitude of DQ 4-Ramsey and STIRAP 4-Ramsey. Amplitude plot of DQ 4-Ramsey and STIRAP 4-Ramsey interferometer techniques. While Fig. 4 shows the accumulated phase ΦΦ\Phiroman_Φ, this plot shows the corresponding amplitude r𝑟ritalic_r (defined in Eq. (9)) for the same data set.

Appendix C Accumulated Phase measurements

As described in Fig. 3, the phase of the final Ramsey pulse is modulated to cancel out the effects of single-quantum coherence, both for DQ 4-Ramsey and STIRAP 4-Ramsey. To describe the process in greater detail, we define the phases of the first Ramsey pulse to be ϕp1subscriptitalic-ϕ𝑝1\phi_{p1}italic_ϕ start_POSTSUBSCRIPT italic_p 1 end_POSTSUBSCRIPT for pump and ϕs1subscriptitalic-ϕ𝑠1\phi_{s1}italic_ϕ start_POSTSUBSCRIPT italic_s 1 end_POSTSUBSCRIPT for Stokes, and we define the phases of the second Ramsey pulse to be ϕp2subscriptitalic-ϕ𝑝2\phi_{p2}italic_ϕ start_POSTSUBSCRIPT italic_p 2 end_POSTSUBSCRIPT for pump and ϕs2subscriptitalic-ϕ𝑠2\phi_{s2}italic_ϕ start_POSTSUBSCRIPT italic_s 2 end_POSTSUBSCRIPT for Stokes. In the non-rotating frame, pulses are then constructed according to Ω~(t)=Real[Ω(t)ei(ϕ+ωt)]~Ω𝑡Realdelimited-[]Ω𝑡superscript𝑒𝑖italic-ϕ𝜔𝑡\widetilde{\Omega}(t)=\mathrm{Real}\left[\Omega(t)e^{i(\phi+\omega t)}\right]over~ start_ARG roman_Ω end_ARG ( italic_t ) = roman_Real [ roman_Ω ( italic_t ) italic_e start_POSTSUPERSCRIPT italic_i ( italic_ϕ + italic_ω italic_t ) end_POSTSUPERSCRIPT ], where ω𝜔\omegaitalic_ω and ϕitalic-ϕ\phiitalic_ϕ are the respective frequency and phase of the RF pulse.

The phases of each pulse in DQ Ramsey are shown in the following table

{tblr}

colspec=—c —— c c c c — c c c c ——, colsep = 4pt \SetCell[c=4]I \SetCell[c=4] Q

R1subscript𝑅1R_{1}italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT R2subscript𝑅2R_{2}italic_R start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT R3subscript𝑅3R_{3}italic_R start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT R4subscript𝑅4R_{4}italic_R start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT R5subscript𝑅5R_{5}italic_R start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT R6subscript𝑅6R_{6}italic_R start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT R7subscript𝑅7R_{7}italic_R start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT R8subscript𝑅8R_{8}italic_R start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT

ϕp1subscriptitalic-ϕ𝑝1\phi_{p1}italic_ϕ start_POSTSUBSCRIPT italic_p 1 end_POSTSUBSCRIPT 0 0 0 0 0 0 0 0

ϕs1subscriptitalic-ϕ𝑠1\phi_{s1}italic_ϕ start_POSTSUBSCRIPT italic_s 1 end_POSTSUBSCRIPT 0 0 0 0 0 0 0 0

ϕp2subscriptitalic-ϕ𝑝2\phi_{p2}italic_ϕ start_POSTSUBSCRIPT italic_p 2 end_POSTSUBSCRIPT 00 π𝜋\piitalic_π π𝜋\piitalic_π 00 0+π40𝜋40+\frac{\pi}{4}0 + divide start_ARG italic_π end_ARG start_ARG 4 end_ARG π+π4𝜋𝜋4\pi+\frac{\pi}{4}italic_π + divide start_ARG italic_π end_ARG start_ARG 4 end_ARG π+π4𝜋𝜋4\pi+\frac{\pi}{4}italic_π + divide start_ARG italic_π end_ARG start_ARG 4 end_ARG 0+π40𝜋40+\frac{\pi}{4}0 + divide start_ARG italic_π end_ARG start_ARG 4 end_ARG

ϕs2subscriptitalic-ϕ𝑠2\phi_{s2}italic_ϕ start_POSTSUBSCRIPT italic_s 2 end_POSTSUBSCRIPT 00 00 π𝜋\piitalic_π π𝜋\piitalic_π 0π40𝜋40-\frac{\pi}{4}0 - divide start_ARG italic_π end_ARG start_ARG 4 end_ARG 0π40𝜋40-\frac{\pi}{4}0 - divide start_ARG italic_π end_ARG start_ARG 4 end_ARG ππ4𝜋𝜋4\pi-\frac{\pi}{4}italic_π - divide start_ARG italic_π end_ARG start_ARG 4 end_ARG ππ4𝜋𝜋4\pi-\frac{\pi}{4}italic_π - divide start_ARG italic_π end_ARG start_ARG 4 end_ARG,

where R14subscript𝑅14R_{1-4}italic_R start_POSTSUBSCRIPT 1 - 4 end_POSTSUBSCRIPT are used to measure I𝐼Iitalic_I, and R58subscript𝑅58R_{5-8}italic_R start_POSTSUBSCRIPT 5 - 8 end_POSTSUBSCRIPT are used to measure Q𝑄Qitalic_Q, according to the equations

I=R1R2+R3R4,Q=R5R6+R7R8.formulae-sequence𝐼subscript𝑅1subscript𝑅2subscript𝑅3subscript𝑅4𝑄subscript𝑅5subscript𝑅6subscript𝑅7subscript𝑅8\displaystyle\begin{split}I&=R_{1}-R_{2}+R_{3}-R_{4}\,,\\ Q&=R_{5}-R_{6}+R_{7}-R_{8}\,.\end{split}start_ROW start_CELL italic_I end_CELL start_CELL = italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL italic_Q end_CELL start_CELL = italic_R start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT . end_CELL end_ROW (8)

The opposite ±π/4plus-or-minus𝜋4\pm\pi/4± italic_π / 4 shifts in the second Ramsey pulse correspond to a π/2𝜋2\pi/2italic_π / 2 phase shift in the effective DQ phase, ϕpϕssubscriptitalic-ϕ𝑝subscriptitalic-ϕ𝑠\phi_{p}-\phi_{s}italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - italic_ϕ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT.

For STIRAP Ramsey, the procedure is identical except S18subscript𝑆18S_{1-8}italic_S start_POSTSUBSCRIPT 1 - 8 end_POSTSUBSCRIPT are used in place of R18subscript𝑅18R_{1-8}italic_R start_POSTSUBSCRIPT 1 - 8 end_POSTSUBSCRIPT. For both DQ Ramsey and STIRAP Ramsey, the accumulated phase ΦΦ\Phiroman_Φ (and amplitude r𝑟ritalic_r, shown in Fig. 6) is obtained by combining I𝐼Iitalic_I and Q𝑄Qitalic_Q into a complex number, according to

z=I+iQ=reiΦ.𝑧𝐼𝑖𝑄𝑟superscript𝑒𝑖Φz=I+iQ=re^{i\Phi}\,.italic_z = italic_I + italic_i italic_Q = italic_r italic_e start_POSTSUPERSCRIPT italic_i roman_Φ end_POSTSUPERSCRIPT . (9)
Refer to caption
Figure 7: Experimentally measured distribution of Rabi frequencies. The distribution of Rabi frequencies was obtained by scanning the duration of a rectangular RF pulse on the f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT transition and performing Fourier analysis on the optically detected signal. The resulting amplitude spectrum is shown for both rectangular and Hann windows. The full width at half maximum (FWHM) was measured to be 0.166Ωe0.166delimited-⟨⟩subscriptΩ𝑒0.166\left<\Omega_{e}\right>0.166 ⟨ roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ⟩.

Appendix D Procedure for Numerical Simulations

Zeeman Shifts. In the simulations, to account for the ac Zeeman shifts, we modify the Hamiltonian in Eq. (4) as

H¯(t)=2(0Ω¯p(t)0Ω¯p(t)2ΔΩ¯s(t)0Ω¯s(t)0),¯𝐻𝑡Planck-constant-over-2-pi20subscript¯Ω𝑝𝑡0subscriptsuperscript¯Ω𝑝𝑡2Δsubscript¯Ω𝑠𝑡0subscriptsuperscript¯Ω𝑠𝑡0\displaystyle\bar{H}(t)=\frac{\hbar}{2}\left(\begin{array}[]{ccc}0&\bar{\Omega% }_{p}(t)&0\\ \bar{\Omega}^{*}_{p}(t)&-2\Delta&\bar{\Omega}_{s}(t)\\ 0&\bar{\Omega}^{*}_{s}(t)&0\end{array}\right)\,,over¯ start_ARG italic_H end_ARG ( italic_t ) = divide start_ARG roman_ℏ end_ARG start_ARG 2 end_ARG ( start_ARRAY start_ROW start_CELL 0 end_CELL start_CELL over¯ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL over¯ start_ARG roman_Ω end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL - 2 roman_Δ end_CELL start_CELL over¯ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL over¯ start_ARG roman_Ω end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL 0 end_CELL end_ROW end_ARRAY ) , (13)

where Ω¯p(t)=Ωp(t)+Ωs(t)eiη(t)subscript¯Ω𝑝𝑡subscriptΩ𝑝𝑡subscriptΩ𝑠𝑡superscript𝑒𝑖𝜂𝑡\bar{\Omega}_{p}(t)=\Omega_{p}(t)+\Omega_{s}(t)e^{-i\eta(t)}over¯ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) = roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) + roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) italic_e start_POSTSUPERSCRIPT - italic_i italic_η ( italic_t ) end_POSTSUPERSCRIPT, Ω¯s(t)=Ωs(t)+Ωp(t)eiη(t)subscript¯Ω𝑠𝑡subscriptΩ𝑠𝑡subscriptΩ𝑝𝑡superscript𝑒𝑖𝜂𝑡\bar{\Omega}_{s}(t)=\Omega_{s}(t)+\Omega_{p}(t)e^{-i\eta(t)}over¯ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) = roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) + roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) italic_e start_POSTSUPERSCRIPT - italic_i italic_η ( italic_t ) end_POSTSUPERSCRIPT, η(t)=(ωpωs)t+ϕpϕs𝜂𝑡subscript𝜔𝑝subscript𝜔𝑠𝑡subscriptitalic-ϕ𝑝subscriptitalic-ϕ𝑠\eta(t)=\left(\omega_{p}-\omega_{s}\right)t+\phi_{p}-\phi_{s}italic_η ( italic_t ) = ( italic_ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) italic_t + italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - italic_ϕ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT with ϕpsubscriptitalic-ϕ𝑝\phi_{p}italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and ϕssubscriptitalic-ϕ𝑠\phi_{s}italic_ϕ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT being the initial phases of the RF pulses. The modified Rabi frequencies take into account that both pump and Stokes fields interact on both transitions in the three-level system. Effectively, this leads to the modulation of the Rabi frequency envelopes with a modulation frequency equal to the difference ωpωssubscript𝜔𝑝subscript𝜔𝑠\omega_{p}-\omega_{s}italic_ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT [41, 42]. A comparison between calculations with and without this correction reveals that the difference is small but noticeable.

State propagation. We obtain the numerical results in Fig. 2 by solving the von Neumann equation. The initial density matrix is a mixed state matching the experimentally measured probabilities at t=0𝑡0t=0italic_t = 0. In Fig. 4 simulation, we assume that the initial state is |1ket1\ket{1}| start_ARG 1 end_ARG ⟩ and solve the Schrödinger equation. In both cases, the condition of two-photon resonance is assumed, δ=ωpωsω1+ω2=0𝛿subscript𝜔𝑝subscript𝜔𝑠subscript𝜔1subscript𝜔20\delta=\omega_{p}-\omega_{s}-\omega_{1}+\omega_{2}=0italic_δ = italic_ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0. The experimental uncertainty associated with the two-photon detuning is approximately 5 Hz. Theoretical modeling and simulations were performed using Julia’s QuantumControl.jl package [43], which uses Chebyshev’s polynomials to calculate the time evolution.

Rabi frequency gradient. We average an ensemble of simulations with different Rabi frequencies to account for RF power gradients across the sensing volume. We take those Rabi frequencies from a Gaussian distribution centered in the reported Rabi frequencies in the main text. The width of the Gaussian distribution, the standard deviation, was verified experimentally and the results are shown in Fig 7. The average population of the state |iket𝑖\ket{i}| start_ARG italic_i end_ARG ⟩ at time t𝑡titalic_t is given by

pi(t)=1σ2πpi(t,Ωe)exp[(ΩeΩe)22σ2]𝑑Ωe.delimited-⟨⟩subscript𝑝𝑖𝑡1𝜎2𝜋superscriptsubscriptsubscript𝑝𝑖𝑡superscriptsubscriptΩ𝑒superscriptsuperscriptsubscriptΩ𝑒delimited-⟨⟩subscriptΩ𝑒22superscript𝜎2differential-dsuperscriptsubscriptΩ𝑒\langle p_{i}(t)\rangle=\frac{1}{\sigma\sqrt{2\pi}}\int\limits_{-\infty}^{% \infty}p_{i}(t,\Omega_{e}^{\prime})\exp\left[-\frac{(\Omega_{e}^{\prime}-% \langle\Omega_{e}\rangle)^{2}}{2\sigma^{2}}\right]d\Omega_{e}^{\prime}\,.⟨ italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) ⟩ = divide start_ARG 1 end_ARG start_ARG italic_σ square-root start_ARG 2 italic_π end_ARG end_ARG ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t , roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) roman_exp [ - divide start_ARG ( roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - ⟨ roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ⟩ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] italic_d roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT . (14)

where Ωedelimited-⟨⟩subscriptΩ𝑒\langle\Omega_{e}\rangle⟨ roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ⟩ is the average peak amplitude of the effective Rabi frequency, σ𝜎\sigmaitalic_σ the standard deviation, and pi(t,Ωe)subscript𝑝𝑖𝑡superscriptsubscriptΩ𝑒p_{i}(t,\Omega_{e}^{\prime})italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t , roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) the population of the state |iket𝑖\ket{i}| start_ARG italic_i end_ARG ⟩ at time t𝑡titalic_t when simulating using a peak amplitude of the effective Rabi frequency ΩesuperscriptsubscriptΩ𝑒\Omega_{e}^{\prime}roman_Ω start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT.

Ramsey signal. The spin-state wavefunction for the DQ Ramsey scheme is given by

|ψf=Uπ/2(2)PτUπ/2(1)PUπ|ψ0ketsubscript𝜓𝑓subscriptsuperscript𝑈2𝜋2subscript𝑃𝜏subscriptsuperscript𝑈1𝜋2𝑃subscript𝑈𝜋ketsubscript𝜓0\ket{\psi_{f}}=U^{(2)}_{\pi/2}P_{\tau}U^{(1)}_{\pi/2}PU_{\pi}\ket{\psi_{0}}| start_ARG italic_ψ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG ⟩ = italic_U start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_π / 2 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_π / 2 end_POSTSUBSCRIPT italic_P italic_U start_POSTSUBSCRIPT italic_π end_POSTSUBSCRIPT | start_ARG italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ (15)

where |ψ0=|0ketsubscript𝜓0ket0\ket{\psi_{0}}=\ket{0}| start_ARG italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ = | start_ARG 0 end_ARG ⟩ is the initial state, Uπsubscript𝑈𝜋U_{\pi}italic_U start_POSTSUBSCRIPT italic_π end_POSTSUBSCRIPT, Uπ/2(1)subscriptsuperscript𝑈1𝜋2U^{(1)}_{\pi/2}italic_U start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_π / 2 end_POSTSUBSCRIPT, Uπ/2(2)subscriptsuperscript𝑈2𝜋2U^{(2)}_{\pi/2}italic_U start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_π / 2 end_POSTSUBSCRIPT are the evolution operators with the Hamiltonian in Eq. (13), the sub-indexes indicate the pulse area. The transformation matrices P𝑃Pitalic_P and Pτsubscript𝑃𝜏P_{\tau}italic_P start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT take into account the phases between pulses and the free-evolution phase accumulated by the states.

For the H-STIRAP Ramsey scheme, |ψ0=|1ketsubscript𝜓0ket1\ket{\psi_{0}}=\ket{1}| start_ARG italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ = | start_ARG 1 end_ARG ⟩, Uπ=P=I^subscript𝑈𝜋𝑃^𝐼U_{\pi}=P=\hat{I}italic_U start_POSTSUBSCRIPT italic_π end_POSTSUBSCRIPT = italic_P = over^ start_ARG italic_I end_ARG, and the evolution operators Uπ/2(1),(2)subscriptsuperscript𝑈12𝜋2U^{(1),(2)}_{\pi/2}italic_U start_POSTSUPERSCRIPT ( 1 ) , ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_π / 2 end_POSTSUBSCRIPT are calculated using pump and Stokes Rabi-frequency envelopes described in the main text, Eq. (7).

The experimentally measured optical readout is the total fluorescence signal from all three nuclear spin states. To account for the relative brightness of the N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N nuclear states, we calculate the Ramsey signal as

R=|1|ψf|2+0.9785|0|ψf|2+0.9861|1|ψf|2,𝑅superscriptinner-product1subscript𝜓𝑓20.9785superscriptinner-product0subscript𝜓𝑓20.9861superscriptinner-product1subscript𝜓𝑓2R=|\langle 1|\psi_{f}\rangle|^{2}+0.9785|\langle 0|\psi_{f}\rangle|^{2}+0.9861% |\langle-1|\psi_{f}\rangle|^{2}\,,italic_R = | ⟨ 1 | italic_ψ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 0.9785 | ⟨ 0 | italic_ψ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 0.9861 | ⟨ - 1 | italic_ψ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (16)

where 1, 0.9785, and 0.9861 are the relative brightness coefficients for the |1ket1\ket{1}| start_ARG 1 end_ARG ⟩, |0ket0\ket{0}| start_ARG 0 end_ARG ⟩, |1ket1\ket{-1}| start_ARG - 1 end_ARG ⟩ states respectively [33].

References

  • Ye and Zoller [2024] J. Ye and P. Zoller, Essay: Quantum sensing with atomic, molecular, and optical platforms for fundamental physics, Phys. Rev. Lett. 132, 190001 (2024).
  • Ludlow et al. [2015] A. D. Ludlow, M. M. Boyd, J. Ye, E. Peik, and P. O. Schmidt, Optical atomic clocks, Rev. Mod. Phys. 87, 637 (2015).
  • Allen and Eberly [1987] L. Allen and J. Eberly, Optical Resonance and Two-level Atoms, Dover books on physics and chemistry (Dover, 1987).
  • Shore [2011] B. W. Shore, Manipulating Quantum Structures Using Laser Pulses (Cambridge University Presss, 2011).
  • Berman and Malinovsky [2011] P. R. Berman and V. S. Malinovsky, Principles of laser spectroscopy and quantum optics (Princeton University Press, 2011).
  • Sola et al. [2018] I. R. Sola, B. Y. Chang, S. A. Malinovskaya, and V. S. Malinovsky, Quantum control in multilevel systems, in Advances In Atomic, Molecular, and Optical Physics, Vol. 67, edited by E. Arimondo, L. F. DiMauro, and S. F. Yelin (Academic Press, 2018) pp. 151–256.
  • Kocharovskaia and Khanin [1986] O. A. Kocharovskaia and I. I. Khanin, Population trapping and coherent transillumination of a three-level medium by a periodic train of ultrashort pulses, Zhurnal Eksperimentalnoi i Teoreticheskoi Fiziki 90, 1610 (1986).
  • Harris [1997] S. E. Harris, Electromagnetically Induced Transparency, Physics Today 50, 36 (1997).
  • Bergmann et al. [1998] K. Bergmann, H. Theuer, and B. W. Shore, Coherent population transfer among quantum states of atoms and molecules, Rev. Mod. Phys. 70, 1003 (1998).
  • Vitanov et al. [2017] N. V. Vitanov, A. A. Rangelov, B. W. Shore, and K. Bergmann, Stimulated raman adiabatic passage in physics, chemistry, and beyond, Rev. Mod. Phys. 89, 015006 (2017).
  • Ni et al. [2008] K.-K. Ni, S. Ospelkaus, M. H. G. de Miranda, A. Pe’er, B. Neyenhuis, J. J. Zirbel, S. Kotochigova, P. S. Julienne, D. S. Jin, and J. Ye, A high phase-space-density gas of polar molecules, Science 322, 231 (2008).
  • Böhm et al. [2021] F. Böhm, N. Nikolay, S. Neinert, C. E. Nebel, and O. Benson, Ground-state microwave-stimulated raman transitions and adiabatic spin transfer in the N15superscriptN15{}^{15}\mathrm{N}start_FLOATSUPERSCRIPT 15 end_FLOATSUPERSCRIPT roman_N nitrogen vacancy center, Phys. Rev. B 104, 035201 (2021).
  • Gong et al. [2024] M. Gong, M. Yu, Y. Chu, W. Chen, Q. Cao, N. Wang, J. Cai, R. Betzholz, and L. Giannelli, Two-photon-transition superadiabatic passage in a nitrogen-vacancy center in diamond, Phys. Rev. A 109, 032626 (2024).
  • Cubel et al. [2005] T. Cubel, B. K. Teo, V. S. Malinovsky, J. R. Guest, A. Reinhard, B. Knuffman, P. R. Berman, and R. G., Coherent population transfer of ground-state atoms into Rydberg states, Phys. Rev. A 72, 023405 (2005).
  • Sola et al. [2022] I. R. Sola, B. Y. Chang, S. A. Malinovskaya, S. C. Carrasco, and V. S. Malinovsky, Stimulated raman adiabatic passage with trains of weak pulses, J. Phys. B: At. Mol. Opt. Phys. 15, 234002 (2022).
  • Bergmann et al. [2019] K. Bergmann, H.-C. Nägerl, C. Panda, G. Gabrielse, E. Miloglyadov, M. Quack, G. Seyfang, G. Wichmann, S. Ospelkaus, A. Kuhn, S. Longhi, A. Szameit, P. Pirro, B. Hillebrands, X.-F. Zhu, J. Zhu, M. Drewsen, W. K. Hensinger, S. Weidt, T. Halfmann, H.-L. Wang, G. S. Paraoanu, N. V. Vitanov, J. Mompart, T. Busch, T. J. Barnum, D. D. Grimes, R. W. Field, M. G. Raizen, E. Narevicius, M. Auzinsh, D. Budker, A. Pálffy, and C. H. Keitel, Roadmap on stirap applications, J. Phys. B: At. Mol. Opt. Phys. 52, 202001 (2019).
  • Malinovsky and Tannor [1997] V. S. Malinovsky and D. J. Tannor, Simple and robust extension of the stimulated raman adiabatic passage technique to N-level systems, Phys. Rev. A 56, 4929 (1997).
  • Vitanov et al. [1999] N. V. Vitanov, K.-A. Suominen, and B. W. Shore, Creation of coherent atomic superpositions by fractional stimulated Raman adiabatic passage, J. Phys. B: At. Mol. Opt. Phys. 32, 4535 (1999).
  • Band and Magnes [1994] Y. B. Band and O. Magnes, Chirped adiabatic passage with temporally delayed pulses, Phys. Rev. A 50, 584 (1994).
  • Chathanathil et al. [2023] J. Chathanathil, A. Ramaswamy, V. S. Malinovsky, D. Budker, and S. A. Malinovskaya, Chirped fractional stimulated raman adiabatic passage, Phys. Rev. A 108, 043710 (2023).
  • Genov et al. [2023] G. T. Genov, S. Rochester, M. Auzinsh, F. Jelezko, and D. Budker, Robust two-state swap by stimulated raman adiabatic passage, J. Phys. B: At. Mol. Opt. Phys. 56, 054001 (2023).
  • Malinovsky and Sola [2004a] V. S. Malinovsky and I. R. Sola, Quantum control of entanglement by phase manipulation of time-delayed pulse sequences. I, Phys. Rev. A 70, 042304 (2004a).
  • Malinovsky and Sola [2004b] V. S. Malinovsky and I. R. Sola, Quantum control of entanglement by phase manipulation of time-delayed pulse sequences. II, Phys. Rev. A 70, 0042305 (2004b).
  • Jain and Kurur [2004] S. Jain and N. D. Kurur, Multiple quantum inversion in scalar coupled systems with amplitude modulated temporally overlapped pulses, Journal of Magnetic Resonance 169, 240 (2004).
  • Hahn [1996] E. L. Hahn, Concepts of nuclear magnetic resonance in quantum optics, in Amazing Light: A Volume Dedicated To Charles Hard Townes On His 80th Birthday, edited by R. Y. Chiao (Springer New York, New York, NY, 1996) pp. 307–323.
  • Jarmola et al. [2021] A. Jarmola, S. Lourette, V. M. Acosta, A. G. Birdwell, P. Blümler, D. Budker, T. Ivanov, and V. S. Malinovsky, Demonstration of diamond nuclear spin gyroscope, Science Advances 7, eabl3840 (2021).
  • Jacques et al. [2009] V. Jacques, P. Neumann, J. Beck, M. Markham, D. Twitchen, J. Meijer, F. Kaiser, G. Balasubramanian, F. Jelezko, and J. Wrachtrup, Dynamic polarization of single nuclear spins by optical pumping of nitrogen-vacancy color centers in diamond at room temperature, Phys. Rev. Lett. 102, 057403 (2009).
  • Smeltzer et al. [2009] B. Smeltzer, J. McIntyre, and L. Childress, Robust control of individual nuclear spins in diamond, Phys. Rev. A 80, 050302 (2009).
  • Steiner et al. [2010] M. Steiner, P. Neumann, J. Beck, F. Jelezko, and J. Wrachtrup, Universal enhancement of the optical readout fidelity of single electron spins at nitrogen-vacancy centers in diamond, Phys. Rev. B 81, 035205 (2010).
  • Fischer et al. [2013] R. Fischer, A. Jarmola, P. Kehayias, and D. Budker, Optical polarization of nuclear ensembles in diamond, Phys. Rev. B 87, 125207 (2013).
  • Barry et al. [2020] J. F. Barry, J. M. Schloss, E. Bauch, M. J. Turner, C. A. Hart, L. M. Pham, and R. L. Walsworth, Sensitivity optimization for nv-diamond magnetometry, Rev. Mod. Phys. 92, 015004 (2020).
  • Dréau et al. [2011] A. Dréau, M. Lesik, L. Rondin, P. Spinicelli, O. Arcizet, J.-F. Roch, and V. Jacques, Avoiding power broadening in optically detected magnetic resonance of single nv defects for enhanced dc magnetic field sensitivity, Phys. Rev. B 84, 195204 (2011).
  • Jarmola et al. [2020] A. Jarmola, I. Fescenko, V. M. Acosta, M. W. Doherty, F. K. Fatemi, T. Ivanov, D. Budker, and V. S. Malinovsky, Robust optical readout and characterization of nuclear spin transitions in nitrogen-vacancy ensembles in diamond, Phys. Rev. Research 2, 023094 (2020).
  • Lourette et al. [2023] S. Lourette, A. Jarmola, V. M. Acosta, A. G. Birdwell, D. Budker, M. W. Doherty, T. Ivanov, and V. S. Malinovsky, Temperature sensitivity of N14superscriptN14{}^{14}\mathrm{N}start_FLOATSUPERSCRIPT 14 end_FLOATSUPERSCRIPT roman_N-v𝑣vitalic_v and N15superscriptN15{}^{15}\mathrm{N}start_FLOATSUPERSCRIPT 15 end_FLOATSUPERSCRIPT roman_N-v𝑣vitalic_v ground-state manifolds, Phys. Rev. Appl. 19, 064084 (2023).
  • Shore [2017] B. W. Shore, Picturing stimulated Raman adiabatic passage: a STIRAP tutorial, Adv. Opt. Photon. 9, 563 (2017).
  • Soshenko et al. [2021] V. V. Soshenko, S. V. Bolshedvorskii, O. Rubinas, V. N. Sorokin, A. N. Smolyaninov, V. V. Vorobyov, and A. V. Akimov, Nuclear spin gyroscope based on the nitrogen vacancy center in diamond, Phys. Rev. Lett. 126, 197702 (2021).
  • Ajoy and Cappellaro [2012] A. Ajoy and P. Cappellaro, Stable three-axis nuclear-spin gyroscope in diamond, Phys. Rev. A 86, 062104 (2012).
  • Riehle [2003] F. Riehle, Atomic and molecular frequency references, in Frequency Standards (John Wiley & Sons, Ltd, 2003) Chap. 5, pp. 117–165.
  • Arunkumar et al. [2023] N. Arunkumar, K. S. Olsson, J. T. Oon, C. A. Hart, D. B. Bucher, D. R. Glenn, M. D. Lukin, H. Park, D. Ham, and R. L. Walsworth, Quantum logic enhanced sensing in solid-state spin ensembles, Phys. Rev. Lett. 131, 100801 (2023).
  • Berman [1997] P. R. Berman, ed., Atom Interferometry (Academic Press, San Diego, CA, 1997).
  • Malinovsky and Rudin [2012a] V. S. Malinovsky and S. Rudin, Geometric single-qubit gates for an electron spin in a quantum dot, Int. J. Quant. Chem. 112, 3744–3749 (2012a).
  • Malinovsky and Rudin [2012b] V. S. Malinovsky and S. Rudin, Ultrafast control of electron spin in a quantum dot using geometric phase, Sol. St. Electr. 78, 28–33 (2012b).
  • Goerz et al. [2022] M. H. Goerz, S. C. Carrasco, and V. S. Malinovsky, Quantum optimal control via semi-automatic differentiation, Quantum 6, 871 (2022).