Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Light pulse storage in Pr:YSO crystal based on the revival of silenced echo protocol

Open Access Open Access

Abstract

We report on the light pulse storage in Pr3+:Y2SiO5 crystal based on the revival of silenced echo protocol, which has the advantage of being immune from the spontaneous emission noise. We optimized the echo retrieval efficiency of the light pulse by employing complex hyperbolic secant rephasing pulses and by finely tuning the optical depth in the inhomogeneous broadening of the crystal. An echo retrieval efficiency of 24.4% was demonstrated, and an optical coherence time of 34.6 μs was extracted from the measured decay dynamics of the echo retrieval efficiency at a cryogenic temperature of 3.4 K. These results could be useful for potential applications in quantum memory and related applications.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Quantum memories, serving as nodes in quantum communication networks, play a pivotal role in mitigating the exponential decay of light in optical fibers [1]. Rare-earth-ion-doped crystals are commonly used as quantum storage materials due to their longer coherence time and wider inhomogeneous linewidth as compared to alkali metal vapours, making them more suitable for integration. Various rare-earth-ion-doped waveguide structures have been implemented for storing optical pulses [24]. Among them, Pr$^{3+}$ ion is a kind of rare earth ion extensively studied in quantum memories. Compared with Er$^{3+}$ ion, which has an odd number of electrons, Pr$^{3+}$ ion is insensitive to the electron spin magnetic moment, therefore, a long coherence time can be obtained even without the need for an extremely low temperature and a strong external magnetic field. Pr$^{3+}$ ion is also one of the few rare earth ions that have achieved the zero first-order Zeeman effect. When an external magnetic field in a specific direction is applied, Pr$^{3+}$ ions can exhibit magnetic insensitivity to external environmental magnetic fields, thereby showing an ultra-long coherence time [5]. In Pr:${\rm Y_2SiO_5}$ (Pr:YSO) crystal, the light pulse storage time as long as 1 minute was reported [6]. In addition, the wavelength of the stored photon (606 nm) is located within the visible spectral range, and research on quantum sources in visible spectral range is relatively mature [7,8]. Also, the dark count rate of a conventional single-photon detector is relatively low in the visible spectral range, and commercial EMCCD and iCCD can be used with high quantum efficiency, which is suitable for quantum image processing.

Currently, classical light pulses have been stored in Pr$^{3+}$ ions via electromagnetically induced transparency (EIT) and photon-echo-based protocol. The photon-echo-based protocol offers a diverse set of techniques including photon echo, controlled reversible inhomogeneous broadening (CRIB), gradient echo memory (GEM), and atomic frequency comb (AFC). However, limited by the storage mechanism, the efficiency based on the EIT effect in Pr:YSO crystals is usually lower than 1% [6,9,10]. With a complex optical cavity and waveform optimization technology, a storage efficiency of 76.3% has been reported in Pr:YSO crystals [10]. For the case of photon echo, the population inversion caused by $\pi$-pulse leads to spontaneous emission noise in the echo signal. The increased quantum noise renders traditional photon echo not suitable for single-photon storage. Various methods have been proposed to address this issue. For example, complex spectral structures can be prepared to avoid population inversion, including CRIB, GEM, AFC, and so on. Based on the orientation of the applied electric field or magnetic field and the transmission direction of the signal light, CRIB can be categorized into transverse CRIB and GEM. Hedges et al. [11] achieved a storage efficiency of 69% in Pr:YSO crystals using the GEM protocol. The AFC protocol has multi-mode storage capacity with relatively high storage efficiency. Amari et al. [12] achieved a storage efficiency of 35% in Pr:YSO crystals through the AFC protocol. Horvath et al. [13] achieved a storage efficiency of 36% by introducing the Stark effect into the AFC protocol. In addition, Sabooni et al. [14] raised the storage efficiency up to 56% using an optical standing wave cavity. Riedmatten et al. [15,16] realized single-photon storage in Pr:YSO crystals based on the AFC protocol.

One notes that all these methods require meticulous preparation processes of specific spectral structures. The revival of silenced echo (ROSE) scheme was proposed to address this challenge [17]. It utilizes two counter-propagating $\pi$-pulses with respect to the signal light to bring the population inversion back to the ground state. This approach effectively reduces the quantum noise in traditional optical echoes without the need for specific spectral tailoring. Also, the first echo, which is usually excited by the first $\pi$-pulse, does not emit actually due to spatial phase mismatch. Bonarotal et al. [18] achieved a background level of one photon in the echo mode, evenly shared between spontaneous emission and coherent noise. Based on the ROSE scheme, Ma et al. [19] proposed the noiseless photon echo (NLPE) protocol, which incorporates two additional energy levels to suppress the noise by frequency filtering, and achieved a fidelity of $0.952 \pm 0.018$ with a storage efficiency of $10.0 \pm 0.4$%.

As compared to all those protocols, the ROSE protocol is relatively simpler, easier to implement, and has higher storage efficiency while without the need for complex optimization. The ROSE protocol has been experimentally verified in Eu$^{3+}$ [20], Tm$^{3+}$ [18,21], and Er$^{3+}$ [21,22] doped crystals, but not for Pr$^{3+}$ ion. In this article, we realize the ROSE protocol in Pr:YSO crystals. Furthermore, we propose a method to optimize the crystal absorption by tuning the laser frequency in the inhomogeneous broadening, and therefore achieving light pulse storage with an echo retrieval efficiency of $24.4{\%}$ in Pr:YSO crystals. This method provides a simple way to dynamically optimize the light pulse storage efficiency and also suits for other atomic storage systems.

2. Experimental setup and principles

The experimental setup is illustrated schematically in Fig. 1(a). The laser employed in this study was a Matisse 2 DX dye laser from Spectra-Physics Inc. The wavelength of the laser beam was precisely tuned to 605.78 nm, corresponding to the atomic transition $^1$D$_2 \leftrightarrow ^3$H$_4$ of Pr$^{3+}$ ions in YSO crystals, and its linewidth was controlled to be less than 100 kHz through the Pound-Drever-Hall techniques. The laser beam was split into two beams, the signal beam and the control beam, via a beam splitter. Both beams were subject to intensity and frequency modulation through acoustic-optical modulators (AOM, MT110-B50A1-VIS, AA opto-electronic) respectively with a double-pass arrangement. The AOMs were digitally controlled by a FPGA card (PCI-7813R, NI). The signal beam was modulated into a Gaussian pulse with its central frequency corresponding to the $\pm \frac {3}{2}\left (e\right )\leftrightarrow \pm \frac {3}{2}\left (g\right )$ energy level transition frequency of the $^1$D$_2 \leftrightarrow ^3$H$_4$ transition of the Pr$^{3+}$ ion (shown in Fig. 1(b)) in Pr:YSO crystal. Subsequently, it was launched into a single-mode optical fiber via an optical coupler (PAF2-7A, Thorlabs Inc) and guided out through another optical coupler (PAF2P-18A, Thorlabs Inc.) for spatial mode optimization. The Gaussian signal pulse was then focused through a plano-convex lens with a focal length of 10 cm and launched into the Pr:YSO crystal. The optical power and the beam waist of the signal pulse were 78 $\mu$W and 56 $\mu$m, respectively. The control beam was modulated into a complex hyperbolic secant (CHS) light pulse by AOM through the LabVIEW program with an optical power of 9.1 mW. Similar to the signal pulse, the control CHS pulse was also coupled into a single-mode fiber and guided out for spatial mode optimization, and then was reflected by a polarization beam splitter and finally launched into the Pr:YSO crystal, enabling counter-propagation with respect to the signal pulse. The Rabi frequency of the control CHS pulses varied with time and was expressed as $\Omega (t)=\Omega _0~\textrm {sech}\left (\beta \left (t-t_{j}\right )\right )$, where $t_j=t_2$ or $t_3$ was the time sequence of the CHS pulses launched into the Pr:YSO crystal, $1/\beta$ and $\Omega _0$ were the pulse duration and the peak Rabi frequency of the CHS pulses, respectively. The frequency was swept around a central frequency $\omega _0$, also corresponding to the $\pm \frac {3}{2}\left (e\right )\leftrightarrow \pm \frac {3}{2}\left (g\right )$ energy level transition frequency of the $^1$D$_2 \leftrightarrow ^3$H$_4$ transition of the Pr$^{3+}$ ion in YSO crystal, and was expressed as $\omega (t)=\omega _0+\mu \beta ~\textrm {tanh}(\beta (t-t_{j}))$ with $t_j=t_2$ or $t_3$, respectively, where $\mu$ is a parameter related to the pulse width, $2\mu \beta$ is the sweep bandwidth of the CHS pulse. In the experiments, $\mu =1$ and $\beta =0.87$ MHz were employed, therefore, the pulse duration $\beta ^{-1}$ and the sweep bandwidth $2\mu \beta$ of the CHS pulses were 1.15 $\mu$s and 1.74 MHz, respectively. Note that we set the parameters of the two CHS pulses the same in the experiments. The CHS $\pi$-pulses are known to exhibit higher population inversion efficiency than other $\pi$ pulses such as square-wave $\pi$-pulses and are widely employed in photon echo experiments [23]. The mode overlap between the signal Gaussian pulse and the control CHS pulse within the crystal could be assessed by checking the reverse-coupling efficiency where the signal pulse was coupled into the single-mode optical fiber of the control CHS pulse. In our experiments, the reverse-coupling efficiency was measured to be 76%. To enhance the signal-to-noise ratio and minimize the back-reflection to the detector, the signal pulse was horizontally polarized and the CHS pulse was vertically polarized by using two half-wave plates (HWPs), as shown in Fig. 1(a).

 figure: Fig. 1.

Fig. 1. (a) Schematic diagram of the experimental setup. QWP: quarter-wave plate. HWP: half-wave plate. Lens: plano-convex lens with a 10-cm focal length. PBS: polarization beam splitter. BS: beam splitter. Coupler: optical fiber coupler. AOM: acousto-optic modulator. PD: photo-detector. The red arrows indicate the propagation directions of the signal pulse and the control CHS pulses, respectively. (b) Energy-level structure of Pr$^{3+}$ ions in Pr:YSO crystal.

Download Full Size | PDF

The Pr:YSO crystal (Physcience Opto-electronics Corp.), with a size of 4 mm $\times$ 5 mm $\times$ 10 mm and a Pr-doping concentration of 0.05%, was placed in a cryostat (Montana Inc.) maintained at a temperature of 3.4 K. The Pr:YSO crystal was of three mutually perpendicular optical axes referred as $\textbf {D}_1$, $\textbf {D}_2$ and b axis, respectively. The $\textbf {D}_1$ axis was along the 4-mm side, the $\textbf {D}_2$ axis was along the 5-mm side, and the b axis was along the 10-mm side, respectively. Both the signal pulse and the control CHS pulses propagated along the b axis. Due to the absence of an anti-reflection coating on the crystal surface, a slight inclination was introduced to eliminate photo-detectors from the stray/reflection lights. In the experiments, we utilized the fine energy levels of the $^1$D$_2 \leftrightarrow ^3$H$_4$ transition of Pr$^{3+}$ ions, which is associated with the $\pm \frac {3}{2}(g)$ ground state and the $\pm \frac {3}{2}(e)$ excited state, as shown in Fig. 1(b), because this energy level transition exhibited the highest absorption. To prepare such a two-level system for storage based on Pr$^{3+}$ ions in YSO crystal, they were initially populated in the $\pm \frac {3}{2}(g)$ ground state through spectral hole burning and anti-hole burning techniques. For details about the spectral hole burning and anti-hole burning techniques, please refer to our previous work in Ref. [25]. It is known that the inhomogeneous broadening of the $^1$D$_2 \leftrightarrow ^3$H$_4$ transition of Pr$^{3+}$ ions in YSO crystal is of the order of 10 GHz in full width at half maximum. The bandwidth of the spectral hole burned on the spectrum of the $^1$D$_2 \leftrightarrow ^3$H$_4$ transition of Pr$^{3+}$ ions in YSO crystal is of tens of MHz at a cryogenic temperature of 3.4 K. While the absorption spectral bandwidth of the transition $\pm \frac {3}{2}\left (e\right )\leftrightarrow \pm \frac {3}{2}\left (g\right )$ is only of hundreds of kHz, much narrower than the bandwidth of the burned spectral hole and the spectral separation between the fine energy levels of $^1$D$_2$ or $^3$H$_4$, as also shown in Fig. 1(b). Therefore, the transition $\pm \frac {3}{2}\left (e\right )\leftrightarrow \pm \frac {3}{2}\left (g\right )$ is spectrally well separated from other transitions such as $\pm \frac {1}{2}\left (e\right )\leftrightarrow \pm \frac {3}{2}\left (g\right )$ and $\pm \frac {5}{2}\left (e\right )\leftrightarrow \pm \frac {3}{2}\left (g\right )$ within the burned spectral hole [25], which justifies the appropriateness of the employed two-level system for storage in Pr:YSO crystal. In contrast to Er$^{3+}$ ions, the fine-structure interval of the $^1$D$_2 \leftrightarrow ^3$H$_4$ transition of Pr$^{3+}$ ions was of a few MHz, much smaller than the inhomogeneous broadening of Pr$^{3+}$ ions. As a result, the energy level preparation plays an important role in the storage. Without prior energy level preparation and when multiple storage processes are executed in rapid succession, the ground-state ions will be gradually depleted, leading to a gradual reduction in the absorption of the signal pulse, and consequently a reduction of the storage efficiency over time. Energy level preparation conducted in advance ensures population stabilization, thereby preserving the stability of storage efficiency.

In the light pulse storage and echo retrieval experiments, a signal $\pi /2$-pulse with a wave vector $\textbf {k}_1$ was launched into the crystal at time $t_1$. Then at time $t_2$, the first control CHS $\pi$-pulse with a wave vector $\textbf {k}_2$ was launched into the crystal in the opposite direction with respect to the signal pulse. However, due to the spatial phase mismatch, no echo was generated even at the time $t_1+2t_{12}$ (where $t_{12}=t_2-t_1$) though the coherence rephases. Nevertheless, due to the first control CHS pulse, the population on the ground state $\pm 3/2(g)$ was transferred to the excited state $\pm 3/2(e)$. At time $t_3$, a second CHS $\pi$-pulse with a wave vector $\textbf {k}_3$ was injected into the crystal in the same direction of the first CHS pulse, and the coherence rephases again at time $t_e=t_1+2t_{23}$ (where $t_{23}=t_3-t_2$). At this point, the spatial phase matching condition was $\textbf {k}_e=2\textbf {k}_3-2\textbf {k}_2+\textbf {k}_1$. Note that the two control CHS pulses were transmitted in the same direction and with the same wave vector, i.e., $\textbf {k}_2=\textbf {k}_3$. Therefore, at time $t_e$, an echo will be emitted from the ground state $\pm 3/2(g)$ with a wave vector $\textbf {k}_e=\textbf {k}_1$ which satisfied the phase matching condition. Moreover, such echo generated from the ground state is immune from the influence of spontaneous emission noise. In addition, the condition $\textbf {k}_e=\textbf {k}_1$ shows that the echo pulse and the signal pulse were propagated in the same direction, therefore, reducing the background noise originated from the control CHS pulse. The light pulse storage time for the entire ROSE process, denoted as $t_{\textrm{stor}}=t_e-t_1$, equals twice the temporal interval $t_{23}=t_3-t_2$. For more comprehensive information, please refer to the content in Appendix A.

3. Experimental results

Figure 2 depict the typical temporal sequence of the signal pulse, the control CHS pulses and the echo pulse respectively in the light pulse storage and echo retrieval process based on the ROSE protocol. Here, the laser frequency was tuned to be of 494736 GHz, i.e., at 605.78 nm. The pulse denoted as S in Fig. 2(a) represented a Gaussian signal pulse, which was launched at time $t_1=2.75~\mu \textrm {s}$. The first control CHS pulse, denoted as CHS$_1$ in Fig. 2(b) was launched at time $t_2=5.07~\mu \textrm {s}$. The second control CHS pulse CHS$_2$ was injected at time $t_3=10.44~\mu \textrm {s}$. Both control CHS pulses CHS$_1$ and CHS$_2$ were of identical parameters with $\mu =1$ and $\beta =0.87~\textrm {MHz}$. The pulse duration $1/\beta$ and the bandwidth $2\mu \beta$ of the control CHS pulses were 1.15 $\mu \textrm {s}$ and 1.74 MHz, respectively. Correspondingly, the echo R was generated at $t_e=13.48~\mu \textrm {s}$. One has to note that, during the AOM modulation of the signal pulse S, the steady-state condition was not satisfied on the rising and falling edges. As a result, there may be a slight discrepancy between the actual frequency and the set frequency of the signal pulse, leading to a slightly low absorption of the signal pulse during these transient phases. This can also be seen from Fig. 2(a) that the signal pulse S exhibits an intensity profile with a shallow valley in the center, indicating that the signal pulse was not in a perfect Gaussian pulse shape.

 figure: Fig. 2.

Fig. 2. (a) The temporal sequence of the signal pulse S and the echo pulse R. (b) The temporal sequence of the control CHS pulses.

Download Full Size | PDF

Theoretically, the echo retrieval efficiency $\eta$ decays exponentially and follows (see Appendix A)

$$\eta={\xi^2_{CHS}}{(\alpha L)}^2e^{-\alpha L}e^{{-}4\gamma t_{23}}$$
where $\xi _{CHS}$ is a correction factor due to the imperfection of the control CHS pulse, $\alpha L$ denotes the optical depth of the crystal with $L$ being the crystal length, and $\gamma$ denotes the decoherence rate of atomic transitions. Figure 3 shows the measured decay dynamics of the echo retrieval efficiency $\eta$ with different temporal interval $t_{12}$ between the first control CHS pulse CHS$_1$ and the signal pulse S. The solid squares are the experimental results, and the red curves are the theoretical fits. In the experiment, for each curve in Fig. 3, we kept the time interval $t_{12}$ the same while adjusting the time interval $t_{23}$ between the first control CHS pulse CHS$_1$ and the second control CHS pulse CHS$_2$ to modulate the storage time $t_{\textrm{stor}}=2t_{23}$. As expected, the echo retrieval efficiency $\eta$ decays exponentially. The echo retrieval efficiency $\eta$ was measured to be $24.4\pm 2.7{\%}$ after a storage time of 11 $\mu$s, indicating an initial storage efficiency of $\eta (t_{\textrm{stor}}=0)=46.5{\%}$. One notes that the storage time $t_{\textrm{stor}}$ can also be expressed as the sum of pulse intervals, i.e. $t_{\textrm{stor}}=t_{12}+t_{23}+(t_e-t_3)=t_{12}+(t_e-t_2)>t_{12}$, indicating that the initial decay with time shorter than $t_{12}$ is not detectable experimentally. Also, by fitting exponentially the measured data using Eq. (1), one gets an optical coherence time of $T_2=\gamma ^{-1}=34.6\, \mu$s. Moreover, one notes from Fig. 3 that, with the same storage time, the measured echo retrieval efficiency was almost the same even with different temporal interval $t_{12}$. This means that the temporal interval $t_{12}$ has little effect on the echo retrieval efficiency, which is consistent with Eq. (1). This result demonstrates the robustness of the ROSE protocol in optical light pulse storage. When the temporal interval $t_{23}$ between the pulses CHS$_1$ and CHS$_2$ are kept constant, the echo, irrespective of the time $t_1$ for the signal pulse launching into the crystal, consistently maintains nearly the same retrieval efficiency. Moreover, the characteristics that the light pulse storage time always equals $2t_{23}$, twice the time interval between pulses CHS$_1$ and CHS$_2$, holds substantial potential applications in areas such as repetition rate extension for pulsed lasers and precision time interval measurements.

 figure: Fig. 3.

Fig. 3. The dependence of the echo retrieval efficiency $\eta$ on the storage time $t_{\textrm{stor}}$. Here $t_{12}$ is the time interval between the signal pulse S and the first CHS pulse CHS$_1$. The solid squares are experimental results, and the red curves are the exponential fits.

Download Full Size | PDF

As shown by Eq. (1), the echo retrieval efficiency of the ROSE protocol depends on the optical depth of the crystal, therefore, it is possible to improve the echo retrieval efficiency by optimizing the optical depth of the crystal. One can easily get from Eq. (1) that the echo retrieval efficiency is optimized when $\alpha L=2$. In general, there are two primary methods to modulate the optical depth of a crystal. One approach involves altering the impurity doping concentration (here Pr-doping concentration) or the length of the crystal. However, when the doping concentration in crystal reaches a certain value, the interaction between impurity ions may only affect the inhomogeneous broadening but has little effect on the absorption of the crystal [24]. In addition, increasing the doping concentration will also reduce the decoherence time [24]. On the other hand, changing the crystal length is also a frequently employed method. Nevertheless, the length of the crystal is usually determined during the design and is not readily adjustable. As an alternative, one can build an optical cavity to enhance the interaction between the light field and the crystal. The second method to adjust the absorption of the crystal is by applying an external magnetic field [17]. However, it requires complex strong magnetic devices positioned around the cryostat.

Here, we proposed an alternative approach to modify the absorption of the crystal within the inhomogeneous broadening of the crystal. In the experiments, we modified the optical depth of the crystal by finely tuning the laser frequency within a narrow spectral range. Here, the laser frequency referred to the central frequency of the signal pulse and the control CHS pulse. The measured dependences of the crystal optical depth and the echo retrieval efficiency on the laser frequency detuning $\Delta \nu$ (with respect to a laser frequency $\nu _0=494740$ GHz) are shown in Fig. 4, where the storage time was set to be 15.6 $\mu$s and keeping other parameters the same. One sees that, within a 20-GHz frequency scanning range, the optical depth of the crystal gradually decreases with the increase of the laser frequency, as indicated by the red squares and red curve in Fig. 4. Correspondingly, the echo retrieval efficiency exhibits an initial increase up to a maximum and then followed by a decrease, as shown by the blue dots and blue curve in Fig. 4. The echo retrieval efficiency reaches the highest when the optical depth is approximately 2, in good agreement with the theoretical prediction [22]. The measured lowest optical depth in Fig. 4 was 0.05, at which the echo retrieval efficiency was about 3.2%. The above result shows that, by fine-tuning the laser frequency, the optical depth of the crystal can be dynamically modified, and consequently the echo retrieval efficiency can be optimized. In addition, the storage bandwidth is related to the frequency scanning range of the control CHS pulse. By increasing the frequency scanning range of the control CHS pulse, the storage bandwidth can be improved. However, it’s essential to adjust the Rabi frequency of the control CHS pulse accordingly to maintain the adiabatic condition.

 figure: Fig. 4.

Fig. 4. The dependence of the echo retrieval efficiency $\eta$ and the crystal optical depth $\alpha L$ on the frequency detuning $\Delta \nu$ with respect to the laser frequency $\nu _0= 494740$ GHz. The storage time was set to be 15.6 $\mu$s in the experiments. The blue dots are the experimental results of the echo retrieval efficiency, which corresponds to the left vertical axis. The red squares are the experimental results for the optical depth, which corresponds to the right vertical axis. The blue and red curves are a guide to the eye.

Download Full Size | PDF

4. Conclusion

In summary, we demonstrated the light pulse storage in Pr:YSO crystal based on the ROSE protocol, which offers the advantages of being immune from the spontaneous emission noise and easy separation of the echo pulse from the rephasing pulses. The echo retrieval efficiency of the light pulse was optimized by employing the CHS rephasing pulses and by optimizing the optical depth of the Pr:YSO crystal within the inhomogeneous broadening through finely tuning the frequency of the interacting light field. An echo retrieval efficiency as high as 24.4% and an optical coherence time of 34.6 ${\rm \mu} s$ were obtained at a cryogenic temperature of 3.4 K. In addition, both theoretical analysis and experimental results show the robustness of ROSE protocol, and the echo retrieval efficiency is predominantly determined by the time interval between the two rephasing CHS pulses. These results provide substantial promise for light pulse storage in Pr$^{3+}$-doped crystals, signifying its importance in quantum memories and related applications.

Appendix A

For a two-level system, when the light field is resonant with the atomic transition, the Schrödinger equation can be written in the interaction picture as

$$i \hbar \frac{d}{d t}\left|\psi_I\right\rangle=\mathcal{V}_j\left|\psi_I\right\rangle \, .$$

Under the rotating wave approximation, the interaction Hamiltonian operator can be written as

$$\mathcal{V}_j=\frac{\hbar \Delta_j}{2} {\sigma}_z^j+\frac{\hbar \Omega}{2}\left( e^{{-}i \boldsymbol{k\cdot r_j}} {\sigma}_{-}^j+ e^{i \boldsymbol{k\cdot r_j}}{\sigma}_{+}^j\right),$$
where $\left |\psi _I\right \rangle$ is the atomic wave function, $\Delta _j$ represents the frequency detuning between the atomic transition of the $j$th atom and the light field, $\Omega$ is the Rabi frequency, ${\sigma }_z$, ${\sigma }_{+}$ and ${\sigma }_{-}$ are the Pauli operators, $\boldsymbol {k}$ is the wave vector of the light field and the subscription $j$ refers to the $j$th atom. To simplify the notation, we omit the subscription $j$ in the following by assuming that all atoms are identical. Assuming that the probability amplitudes of the atoms in the ground state and the excited state are $c_b$ and $c_a$, and by substituting them into Eq. (A1), one gets
$$\left(\begin{array}{l} c_a(t) \\ c_b(t) \end{array}\right)=\left(\begin{array}{cc} \cos \left(\frac{\Omega_g \mathrm{t}}{2}\right)-i \frac{\Delta}{\Omega_g} \sin \left(\frac{\Omega_g \mathrm{t}}{2}\right) & i e^{{-}i \boldsymbol{k\cdot r}} \frac{\Omega}{\Omega_g} \sin \left(\frac{\Omega_g \mathrm{t}}{2}\right) \\ i e^{i \boldsymbol{k\cdot r}} \frac{\Omega}{\Omega_g} \sin \left(\frac{\Omega_g \mathrm{t}}{2}\right) & \cos \left(\frac{\Omega_g \mathrm{t}}{2}\right)+i \frac{\Delta}{\Omega_g} \sin \left(\frac{\Omega_g \mathrm{t}}{2}\right) \end{array}\right)\left(\begin{array}{c} c_a(0) \\ c_b(0) \end{array}\right),$$
where $\Omega _g=\sqrt {\Omega ^{2}+\Delta ^{2}}$. The evolution matrix of the probability amplitude can be divided into two cases. The first case satisfies the condition $\Omega >>\Delta$, where the external field plays a major role. The evolution matrix in this case is defined as $\mathcal {T}(\Theta )$, which represents the influence of the external field on the atomic state. Here, $\Theta =\Omega t$ represents the pulse area of the external field. The second one is for the case $\Omega <<\Delta$, the change in the atomic state is mainly influenced by the frequency detuning $\Delta$. The transformation matrix in this case is defined as $\mathcal {U}(t)$, where $t$ is the evolution time. We will focus on the derivations of the corresponding transformation matrices $\mathcal {T}(\Theta )$ and $\mathcal {U}(t)$ for these two cases.

When the external field is strong enough so that the evolution caused by the frequency detuning is negligible, i.e. $\Omega >>\Delta$, Eq. (A3) can be simplified as

$$\left(\begin{array}{c} c_a(t) \\ c_b(t) \end{array}\right)=\left(\begin{array}{cc} \cos (\Theta / 2) & i e^{{-}i\boldsymbol{k\cdot r}} \sin (\Theta / 2) \\ i e^{i \boldsymbol{k\cdot r}} \sin (\Theta / 2) & \cos (\Theta / 2) \end{array}\right)\left(\begin{array}{l} c_a(0) \\ c_b(0) \end{array}\right)\, .$$

In experiments, the light wave propagates along the $z$-axis perpendicular to the surface of the crystal, so $\boldsymbol {k\cdot r}=kz$. For the signal pulse, one sets $\Theta =\pi /2$, therefore, one gets

$$\mathcal{T}_{\pi/2}=\frac{\sqrt{2}}{2}\left(\begin{array}{cc} 1 & ie^{- i k z} \\ ie^{ i k z} & 1 \end{array}\right)\, .$$

While for the CHS pulse, one sets $\Theta =\pi$ for simplicity, and gets

$$\mathcal{T}_{\pi}=\left(\begin{array}{cc} 0 & ie^{- i k z} \\ ie^{ i k z} & 0 \end{array}\right)\, .$$

When there is no external field acting on the atom, the change in the atomic state is mainly caused by the frequency detuning $\Delta$. In this case, the probability amplitudes of the atoms being in the ground or excited states change with time and are given by

$$\left(\begin{array}{c} c_a(t) \\ c_b(t) \end{array}\right)=\left(\begin{array}{cc} e^{i \Delta t/2 } & 0 \\ 0 & e^{{-}i \Delta t/2} \end{array}\right)\left(\begin{array}{l} c_a(0) \\ c_b(0) \end{array}\right)\, .$$

Therefore, $\mathcal {U}(t)$ can be written as

$$\mathcal{U}(t)=\left(\begin{array}{cc} e^{i \Delta t/2 } & 0 \\ 0 & e^{{-}i \Delta t/2} \end{array}\right)\, .$$

Let us consider the light pulse storage and echo retrieval process based on the ROSE protocol. The two-level atomic ensemble is initially populated on the ground state $\left |\psi _I(0)\right \rangle$ with $c_a(0)=0$ and $c_b(0)=1$, and a signal $\pi /2$-pulse is launched into the crystal and interacts with the two-level atomic ensemble at time $t_1$, therefore the atomic ensemble evolves into the state $\left |\psi _I(t_1)\right \rangle =\mathcal {T}_{\pi /2} \left |\psi _I(0)\right \rangle$. Then the atomic ensemble evolves in dark, and at time $t_2$ the first control CHS $\pi$-pulse is launched into the crystal in the opposite direction with respect to the signal pulse, and the atomic ensemble evolves into the state $\left |\psi _I(t_2)\right \rangle =\mathcal {T}_{\pi } \mathcal {U}(t_2-t_1) \mathcal {T}_{\pi /2} \left |\psi _I(0)\right \rangle$. The second control CHS $\pi$-pulse is launched into the crystal at time $t_3$, and the echo emits at time $t_e$. At this moment, the atomic ensemble state can be written as $\left |\psi _I(t_e)\right \rangle =\mathcal {U}(t_e-t_3) \mathcal {T}_{\pi } \mathcal {U}(t_3-t_2) \mathcal {T}_{\pi } \mathcal {U}(t_2-t_1) \mathcal {T}_{\pi /2} \left |\psi _I(0)\right \rangle$. Also, one has to consider the atomic decoherence effect with a decoherence rate $\gamma$. Therefore, according to Eq. (A5), Eq. (A6) and Eq. (A8), one can deduce the density matrix element $\rho _{ba}(t_e)$ of the atomic transition at the echo emitting time $t_e=2t_{23}+t_1$

$$\begin{aligned} \rho_{ba}(t_e) &= c_b(t_e)c_a(t_e)e^{-\gamma(t_e-t_1)} \\ &= e^{i(2k_3-2k_2+k_1)z}e^{i\Delta({-}t_{12}+t_{23}-t_e-t_3)}e^{-\gamma(t_e-t_1)} \, , \end{aligned}$$
where $t_{nm}=t_m-t_n\,(m,\, n=1,2,3)$.

By substituting $\rho _{b a}(t_e)$ into the Helmholtz equation, one can get the motion equation of the rephasing echo at time $t_e$,

$$\frac{\partial \mathcal{E}}{\partial z} + \frac{\alpha}{2}\mathcal{E}=ig^2 N e^{-\alpha z} e^{i(2k_3-2k_2+k_1)z} e^{-\gamma(t_e-t_1)}e^{{-}ik_ez} \int d \Delta f(\Delta)e^{i\Delta({-}t_{12}+t_{23}-t_e-t_3)}\mathcal{E}_{in},$$
where $\alpha$ is the absorption coefficient of the crystal, $f(\Delta )$ is the density of states and $\mathcal {E}_{in}$ is the signal amplitude at the incident surface of the crystal, $g$ is the coupling constant between the light field and the atoms, $N$ is the number of atoms interacting with the light fields.

When the spatial phase matching condition is satisfied, that is $k_e=2k_3-2k_2+k_1$, and assuming the density of states $f(\Delta )$ obeys the Gaussian distribution, one gets

$$\begin{gathered} \mathcal{E}_{out}={-}\alpha Le^{-\alpha L/2} e^{{-}2\gamma t_{23}}\mathcal{E}_{in} \end{gathered},$$
where $L$ is the crystal thickness and $\alpha L$ represents the optical depth. $\mathcal {E}_{out}=\mathcal {E}(L)$ means the field amplitude of the output echo. Similar result was also obtained by M. Afzelius et al. [26]. Note that, due to the existence of the optical depth, both the signal and the control CHS pulses decrease along the propagation pathes, therefore deviate from the ideal $\pi /2$ or $\pi$ pulses and finally influence the echo retrieval efficiency. For detailed calculation, however, one has to resort to the pulse area theorem [27], which is out of the scope of this paper. Here, for simplicity, we consider these effects by introducing a correction factor $\xi _{CHS}$. Thus, the echo retrieval efficiency $\eta$ based on the ROSE protocal can be expressed as
$$\begin{gathered} \eta=\left|\frac{\mathcal{E}_{out}}{\mathcal{E}_{in}}\right|^2=(\xi_{CHS})^2(\alpha L)^2e^{-\alpha L} e^{{-}4\gamma t_{23}} \end{gathered}\, .$$

Recently, S. A. Moiseev et al. [28] discussed the influence of the optical depth of the signal and control pulses on the echo efficiency based on the ROSE protocol in ${\rm TM}^{3+}$:$\textrm{Y}_{3}\textrm{Al}_{5}\textrm{O}_{12}$ waveguide, and similar tendency as described by Eq. (A12) was obtained.

Funding

National Natural Science Foundation of China (12104242, 91750204); 111 Project (B23045); Fundamental Research Funds for the Central Universities.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data may be obtained from the authors upon reasonable request.

References

1. H.-J. Briegel, W. Dür, J. I. Cirac, et al., “Quantum repeaters: the role of imperfect local operations in quantum communication,” Phys. Rev. Lett. 81(26), 5932–5935 (1998). [CrossRef]  

2. G. Corrielli, A. Seri, M. Mazzera, et al., “Integrated optical memory based on laser-written waveguides,” Phys. Rev. Appl. 5(5), 054013 (2016). [CrossRef]  

3. E. Saglamyurek, M. Grimau Puigibert, Q. Zhou, et al., “A multiplexed light-matter interface for fibre-based quantum networks,” Nat. Commun. 7(1), 11202 (2016). [CrossRef]  

4. C. Liu, T. X. Zhu, M. X. Su, et al., “On-demand quantum storage of photonic qubits in an on-chip waveguide,” Phys. Rev. Lett. 125(26), 260504 (2020). [CrossRef]  

5. M. Hain, M. Stabel, and T. Halfmann, “Few-photon storage on a second timescale by electromagnetically induced transparency in a doped solid,” New J. Phys. 24(2), 023012 (2022). [CrossRef]  

6. G. Heinze, C. Hubrich, and T. Halfmann, “Stopped light and image storage by electromagnetically induced transparency up to the regime of one minute,” Phys. Rev. Lett. 111(3), 033601 (2013). [CrossRef]  

7. J. Fekete, D. Rieländer, M. Cristiani, et al., “Ultranarrow-band photon-pair source compatible with solid state quantum memories and telecommunication networks,” Phys. Rev. Lett. 110(22), 220502 (2013). [CrossRef]  

8. J. J. Liu, J. C. Liu, P. Yu, et al., “Sub-megahertz narrow-band photon pairs at 606 nm for solid-state quantum memories,” APL Photonics 5(6), 066105 (2020). [CrossRef]  

9. J. J. Longdell, E. Fraval, M. J. Sellars, et al., “Stopped light with storage times greater than one second using electromagnetically induced transparency in a solid,” Phys. Rev. Lett. 95(6), 063601 (2005). [CrossRef]  

10. D. Schraft, M. Hain, N. Lorenz, et al., “Stopped light at high storage efficiency in a Pr3+:Y2SiO5 crystal,” Phys. Rev. Lett. 116(7), 073602 (2016). [CrossRef]  

11. M. P. Hedges, J. J. Longdell, Y. Li, et al., “Efficient quantum memory for light,” Nature 465(7301), 1052–1056 (2010). [CrossRef]  

12. A. Amari, A. Walther, M. Sabooni, et al., “Towards an efficient atomic frequency comb quantum memory,” J. Lumin. 130(9), 1579–1585 (2010). [CrossRef]  

13. S. P. Horvath, M. K. Alqedra, A. Kinos, et al., “Noise-free on-demand atomic frequency comb quantum memory,” Phys. Rev. Res. 3(2), 023099 (2021). [CrossRef]  

14. M. Sabooni, Q. Li, S. Kroll, et al., “Efficient quantum memory using a weakly absorbing sample,” Phys. Rev. Lett. 110(13), 133604 (2013). [CrossRef]  

15. D. Rielander, K. Kutluer, P. M. Ledingham, et al., “Quantum storage of heralded single photons in a praseodymium-doped crystal,” Phys. Rev. Lett. 112(4), 040504 (2014). [CrossRef]  

16. A. Seri, D. Lago-Rivera, A. Lenhard, et al., “Quantum storage of frequency-multiplexed heralded single photons,” Phys. Rev. Lett. 123(8), 080502 (2019). [CrossRef]  

17. S. A. Moiseev and S. Kroll, “Complete reconstruction of the quantum state of a single-photon wave packet absorbed by a doppler-broadened transition,” Phys. Rev. Lett. 87(17), 173601 (2001). [CrossRef]  

18. M. Bonarota, J. Dajczgewand, A. Louchet-Chauvet, et al., “Photon echo with a few photons in two-level atoms,” Laser Phys. 24(9), 094003 (2014). [CrossRef]  

19. Y. Z. Ma, M. Jin, D. L. Chen, et al., “Elimination of noise in optically rephased photon echoes,” Nat. Commun. 12(1), 4378 (2021). [CrossRef]  

20. C. Liu, Z. Q. Zhou, T. X. Zhu, et al., “Reliable coherent optical memory based on a laser-written waveguide,” Optica 7(2), 192–197 (2020). [CrossRef]  

21. V. Damon, M. Bonarota, A. Louchet-Chauvet, et al., “Revival of silenced echo and quantum memory for light,” New J. Phys. 13(9), 093031 (2011). [CrossRef]  

22. J. Dajczgewand, J. L. Le Gouet, A. Louchet-Chauvet, et al., “Large efficiency at telecom wavelength for optical quantum memories,” Opt. Lett. 39(9), 2711–2714 (2014). [CrossRef]  

23. F. de Seze, F. Dahes, V. Crozatier, et al., “Coherent driving of Tm3+:YAG ions using a complex hyperbolic secant optical field,” Eur. Phys. J. D 33(3), 343–355 (2005). [CrossRef]  

24. C. Simon, M. Afzelius, J. Appel, et al., “Quantum memories: A review review based on the European integrated project “Qubit Applications (QAP)”,” Eur. Phys. J. D 58(1), 1–22 (2010). [CrossRef]  

25. Y. F. Tu, G. Q. Zhang, Z. H. Zhai, et al., “Angular multiplexing storage of light pulses and addressable optical buffer memory in Pr3+:Y2SiO5 based on electromagnetically induced transparency,” Phys. Rev. A 80(3), 033816 (2009). [CrossRef]  

26. M. Afzelius, C. Simon, H. de Riedmatten, et al., “Multimode quantum memory based on atomic frequency combs,” Phys. Rev. A 79(5), 052329 (2009). [CrossRef]  

27. S. L. McCall and E. L. Hahn, “Self-induced transparency,” Phys. Rev. 183(2), 457–485 (1969). [CrossRef]  

28. S. A. Moiseev, M. M. Minnegaliev, E. S. Moiseev, et al., “Pulse-area theorem in a single-mode waveguide and its application to photon echo and optical memory in Tm3+:Y3Al5O12,” Phys. Rev. A 107(4), 043708 (2023). [CrossRef]  

Data availability

Data may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. (a) Schematic diagram of the experimental setup. QWP: quarter-wave plate. HWP: half-wave plate. Lens: plano-convex lens with a 10-cm focal length. PBS: polarization beam splitter. BS: beam splitter. Coupler: optical fiber coupler. AOM: acousto-optic modulator. PD: photo-detector. The red arrows indicate the propagation directions of the signal pulse and the control CHS pulses, respectively. (b) Energy-level structure of Pr$^{3+}$ ions in Pr:YSO crystal.
Fig. 2.
Fig. 2. (a) The temporal sequence of the signal pulse S and the echo pulse R. (b) The temporal sequence of the control CHS pulses.
Fig. 3.
Fig. 3. The dependence of the echo retrieval efficiency $\eta$ on the storage time $t_{\textrm{stor}}$. Here $t_{12}$ is the time interval between the signal pulse S and the first CHS pulse CHS$_1$. The solid squares are experimental results, and the red curves are the exponential fits.
Fig. 4.
Fig. 4. The dependence of the echo retrieval efficiency $\eta$ and the crystal optical depth $\alpha L$ on the frequency detuning $\Delta \nu$ with respect to the laser frequency $\nu _0= 494740$ GHz. The storage time was set to be 15.6 $\mu$s in the experiments. The blue dots are the experimental results of the echo retrieval efficiency, which corresponds to the left vertical axis. The red squares are the experimental results for the optical depth, which corresponds to the right vertical axis. The blue and red curves are a guide to the eye.

Equations (13)

Equations on this page are rendered with MathJax. Learn more.

η = ξ C H S 2 ( α L ) 2 e α L e 4 γ t 23
i d d t | ψ I = V j | ψ I .
V j = Δ j 2 σ z j + Ω 2 ( e i k r j σ j + e i k r j σ + j ) ,
( c a ( t ) c b ( t ) ) = ( cos ( Ω g t 2 ) i Δ Ω g sin ( Ω g t 2 ) i e i k r Ω Ω g sin ( Ω g t 2 ) i e i k r Ω Ω g sin ( Ω g t 2 ) cos ( Ω g t 2 ) + i Δ Ω g sin ( Ω g t 2 ) ) ( c a ( 0 ) c b ( 0 ) ) ,
( c a ( t ) c b ( t ) ) = ( cos ( Θ / 2 ) i e i k r sin ( Θ / 2 ) i e i k r sin ( Θ / 2 ) cos ( Θ / 2 ) ) ( c a ( 0 ) c b ( 0 ) ) .
T π / 2 = 2 2 ( 1 i e i k z i e i k z 1 ) .
T π = ( 0 i e i k z i e i k z 0 ) .
( c a ( t ) c b ( t ) ) = ( e i Δ t / 2 0 0 e i Δ t / 2 ) ( c a ( 0 ) c b ( 0 ) ) .
U ( t ) = ( e i Δ t / 2 0 0 e i Δ t / 2 ) .
ρ b a ( t e ) = c b ( t e ) c a ( t e ) e γ ( t e t 1 ) = e i ( 2 k 3 2 k 2 + k 1 ) z e i Δ ( t 12 + t 23 t e t 3 ) e γ ( t e t 1 ) ,
E z + α 2 E = i g 2 N e α z e i ( 2 k 3 2 k 2 + k 1 ) z e γ ( t e t 1 ) e i k e z d Δ f ( Δ ) e i Δ ( t 12 + t 23 t e t 3 ) E i n ,
E o u t = α L e α L / 2 e 2 γ t 23 E i n ,
η = | E o u t E i n | 2 = ( ξ C H S ) 2 ( α L ) 2 e α L e 4 γ t 23 .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.