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

Solid-state pulsed microwave emitter based on Rydberg excitons

Open Access Open Access

Abstract

We theoretically demonstrate how the cuprous oxide Cu 2O could be used as a gain medium in a solid-state maser. By taking advantage of radiative microwave transitions between highly excited Rydberg states, one can achieve population inversion and masing in a wide range of wavelengths. In the pulsed emission regime, the considered excitonic system is characterized by intricate and rich dynamics, which are investigated numerically, taking into account several key features of the medium, such as strong Stark shift of energy levels and the presence of the Rydberg blockade effect.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

For many years, excitons have played an important role in the description of optical properties of insulators and semiconductors. They consist of an electron and a hole which are bounded by their Coulomb attraction, closely resembling atoms with a series of energy levels similar to hydrogen. Cuprous oxide Cu 2O is a generic example in which four different excitonic states (yellow, green, blue and violet) have been observed. In 2014 an outstanding experiment realized by the Dortmund group [1] has reneved the interest in the field of excitons as it turned out that the yellow series could be followed up to high principal quantum number n=25. Those high-lying excitons in analogy to Rydberg atoms have been called Rydberg excitons (RE). A large spatial extent of RE reaching micrometers, significantly exceeding the length of the wave which has created them, large lifetimes which scale as n2 and the energy spacing of neighboring states which decreases as n3 allow for an observation of RE in ranges of external parameters much different from other quantum situations. Their excitation energy of about 90 meV is lower by two orders of magnitude than the atomic Rydberg energy. This small binding energy of RE makes them sensitive to external electric or magnetic fields as compared with other systems. Those specific properties od RE in Cu 2O have motivated both theoretical and experimental interest in this field, with studies ranging from their spectroscopy, i.e., optical [2], electrooptical [3] and magnetooptical spectra [4] through non-atomic scaling laws [5] to quantum chaos [6]. Rydberg excitons can easily be excited and manipulated with laser light and they feature a quasicontinuum of narrow states with strong transition coupling to microwave and terahertz radiation. Those two properties endow RE with unusual potential for applications combining optical light with THz waves and microwaves in quantum sensing and information processing.

On the other hand, investigations of the dynamical properties of the systems with RE have recently attracted more interest. The dipole moments of RE, due to a large orbital radius, are exceptionally large for higher values of n. Therefore, the interaction between high lying Rydberg excitons is particularly strong and leads to the so-called Rydberg blockade that prevents the optical excitation of nearby excitons by shifting their corresponding levels out of the resonance with the exciting electromagnetic field. The appearance of Rydberg blockade clearly distinguishes linear and nonlinear regimes of dynamical phenomena regarding RE. In the nonlinear regime, the possibility of observing giant optical nonlinearities of RE in microcavities has been considered [7], while the paper [8] deals with strong interaction phenomena at very low densities of RE, which enables one to determine the contribution of the nonlinear optical response of the medium. In the case of a smaller exciton density it is possible to remain in the linear range for which a single-photon source has been proposed [9].

Recently a lot of effort has been devoted to develop a solid-state maser devices which are characterized by compact size, material stability and high optical pumping efficiency. While the generation of gain based on population inversion due to electromagnetically induced transparency in N-type four-level atomic systems has been dicussed in [10, 11] we focus our attention on RE in couprus oxide. Concerning RE it is worth mentioning that the wavelenghts corresponding to the transitions between higher states fall into sub- and milimeter region. Matching together their attractive properties, i.e., long lifetimes, long dephasing rates, huge dipole moments and their sensitivity to external fields, we have proposed to use an ensemble of optically pumped RE in cuprous oxide as the gain medium for a continuous-wave maser oscillators. The existing atomic gaseous masers demand high vacuum [12] and many solid-state masers require helium temperatures [13]. Moreover, once constructed, such a maser operates at a fixed wavelength. Continuous-wave room temperature maser has been first realized in 2012 in the organic material of p-terpenthenyl molecular crystal [14] and then room-temperature solid-state maser and microwave amplifiers using nitrogen-vacancy NV centere in diamond have been constructed [15, 16].

It seems that Rydberg excitons are promising candidates for the realization of the population inversion, which leads to a tunable maser [17] because the excitonic levels can be accessed separately by tuning the wavelength of the optical excitationvia an external electric field, which due to Stark shift, can displace the resonances (energy levels) [18]. By tuning specific transitions to the frequency of microwave cavity, one can establish population inversion, so that the Cu 2O crystal becomes a gain medium. Although RE were observed at helium temperatures, we claim that in principle a maser might operate up to nitrogen or even higher temperatures [18] and it can be tuned by an external electric field, generating a wide range of emission at different wavelengths with a significant output power. In this paper we theoretically analyse and discuss in detail the case of pulse maser established in a three or four-level system consisting of of Rydberg excitons in Cu 2O. Our numerical simulation demonstrates that masing and microwave amplification are feasible under accessible conditions and our devices can reach output power up to 103 W at the temperature of 100 K.

In the first section we recall the characteristic features of REs particularly useful for obtaining population inversion and stimulated microwave emission. Next, we propose a masing systems based on a ground state and three or two excitonic levels. The basic equations describing the population dynamics are presented and the influence of the microwave cavity geometry and Q factor on the performance is indicated. The following sections 4 and 5 are devoted to the discussion of numerical results obtained for three- and four- level maser systems. The general characteristics of setups based on all accessible state configurations are presented and selected, representative examples are discussed in detail. Finally, the conclusions are presented.

2. Rydberg exciton properties

The Rydberg excitons in Cu 2O feature several unique properties, making this material especially suitable as a gain medium in a solid-state maser. Importantly, they exhibit many similarities to Rydberg atoms, which have been successfully used for stimulated microwave emission in atomic vapor [19]. The key common properties and differences between those media are outlined below.

One of the striking characteristics of Rydberg atoms and Rydberg excitons are their large sizes. Due to the large distance between electron and hole, the P excitons with large n are characterized by enormous dipole moments. Assuming a hydrogen-like wavefunction, the radius of exciton with quantum number n is [1]

rn=12aB[3n2l(l+1)]
where aB is the Bohr radius. Thus, one can expect that the exciton dipole moments range from 1.5eaB (1S state) up to 150eaB (10P state) and even more for larger n. In this paper, we focus only on the S and P excitons and radiative transitions between these states.

Keeping the assumption of hydrogen-like wavefunctions one can calculate from their overlap the SP transition dipole moments dij [17, 20]. Our numerical results for the first 10 states are shown in the Table 2.

Tables Icon

Table 1. Transition dipole moments dn1n2 between n1S (rows) and n2P (columns) states given in units of eaB.

From the calculated values shown in the Table 1, several general observations follow; the largest moments are obtained for nPnS transitions. The magnitude of those moments scales roughly as n2. For n2n1 transition (n2>n1), the scaling is on the order of 4(n1n2).

The large excitonic radius is a limiting factor to the maximum density of excitons. For a sufficient concentration, the efficiency of exciton creation is diminished due to the Rydberg blockade mechanism [1]; as the mean distance between excitons decreases, their mutual dipole-dipole interaction causes a shift of energy levels which prevents further light absorption. In our calculations this mechanism is taken into account by multiplying the absorbed pump power by a factor b=(1ρVblockade) [9], where the Vblockade is the volume occupied by the pumped excitonic state, which is given by

Vblockade=31016n7[mm3].

This estimation can be used to calculate the upper limit of the population; for example, for low n = 3 state, assuming that the whole available volume is occupied by excitons, one can achieve exciton density of up to ρ=1.51012 mm 3. This value is much larger than the density of Rydberg atoms used in masers [19] and comparable with other solid-state devices such as the diamond maser [16], which is based on N1014 active NV centres.

As shown by Kazimierczuk et al. [1], by increasing the input power from P=0.01 mW/mm 2 to P = 1 mW/mm 2, one can expect a 90% reduction of absorption for n=18 state and 99% reduction for n=22. On the other hand, due to the fact that the blockade efficiency scales with the state number as n10, it is dramatically smaller for low n states; for n<10 the absorption efficiency should remain high even for input power of up to 100 W/mm 2, where thermal considerations become the limiting factor.

Another characteristic aspect of Rydberg excitons, which will be usefull for obtaining population inversion, are their unusually long lifetime and the corresponding dissipation parameters γ. For the P exciton with quantum number n, one can use a fit [2] to the experimental data [1]

γ0P(n)=24 meV1+0.01n2n3
which follows the standard n3 dependence of the damping factor on the state number and includes the observed deviation from this law for n>10; γ0P is the damping rate without phonon interaction. While REs are observed at cryogenic conditions, one can in principle detect them at higher temperatures [21]. The temperature dependence of damping according to [22, 23] is described by the following expression
γP(n)=γ0P(n)+γACT+γLO[exp (ωLO/kBT)1]1.

The second term of the r.h.s. of above formula represents exciton scattering on acoustic (AC) phonons, and the third one is due to interactions with longitudinal-optical (LO) phonons. The coefficients γAC=22.9103 meV/K and γLO=1633 meV represent the magnitude of those two interactions, respectively. The relation predicts almost a constant linewidth below T=50 K, as explained by Stolz et al [24]. This effect has been taken into account in other models [25] and has also been observed experimentally by Kitamura et al. [21]. For the non-radiative relaxation of S - excitons, one can use a general estimation γS0.1γP [26]. Finally, since the transition probability between states i,j scales as ωij3 [5, 27], one can estimate the inter-excitonic transition rates

γijγiωij3ωi3.
where ωij is in the microwave range and ωi is the optical frequency of photon needed to create the exciton.

The energy levels of the upper excitonic states are located closely together, so that the inter-excitonic transitions are located in the microwave regime. The exciton energy depends on the quantum number n as follows

E=EgRy(nδp)2,
where Eg=2172.08 meV, Ry=92 meV is the Rydberg energy, δp=0.23 is the quantum defect originating from Cu 2O band structure [1, 28]. As mentioned before, when an electric field is applied, the high n levels are subject to significant Stark shifts. Basing on the real density matrix approach presented in [3], for the applied field F [V/cm], one can develop an approximation for the energy shift
ΔE=2.4106n[4.2(n8.452)2+(n35.16)3]F(1+2.2/n)[eV],
where the parameters have been obtained from the fitting [18] and predicted shifts are consistent with experimental observations [29].

3. System setup

As in our previous work [17], the system consists of a Cu 2O crystal placed inside a metal cavity (See Fig. 1, left panel). The cavity is characterized by a resonant frequency ωc2 and a quality factor of Q2=105, which is a typical of dielectric-loaded cavities [16, 30] and can reach values of up to 108 for superconductor cavities [31]. The crystal has a form of a cylinder bounded by two parallel metallic mirrors (one semi-transparent), forming an inner cavity with frequency ωc1 and Q1=104. These two frequencies are tuned to the respective transitions in a 4-level system based on excitonic levels (Fig. 1, right panel). Due to the Purcell effect, which is an environment-induced enhancement of the rate of spontaneous emission [17, 32], the probability of the selected microwave transition ij is muliplied by a factor proportional to the Q factor and a ratio λij3/V, where V is a value close to the geometric volume of the cavity [33], which is in our case of order of millimeters. The purpose of the outer cavity is to promote a fast transition between n3P and n2S state, leaving the upper, optically pumped level relatively empty. The n2 level, which due to selections rules has to be S-exciton state, is relatively metastable; its population increases up to the point where the masing action starts based on n2Sn1P transition. To describe the dynamics of the system, we used a modified set of the rate equations [17, 34]

N3±it=bPrγ3±iN3±iγ3±i,2N3±iN2t=γ3±i,2N3±iiγ2N2γ2,1N2(N2N1)BWN1t=γ2,1N2+(N2N1)BWγ1N1Wt=ω(N2N1)BW2γc1W
where Ni is the population of the ni level. Note that the rate equations follow from the general Bloch equations for the density matrix after adiabatic elimination of non-diagonal elements of the latter has been performed. Such an approximation is justified if relaxation and dephasing rates are greater then the spectral width of the pulse [35]. We have verified that these conditions are fulfilled in the following discussion. The electromagnetic field density in the cavity is denoted by W and B=π3ϵ02|d12|2P12 is the Einstein coefficient, which depends on the transition dipole moment d12 and also includes the cavity-dependent Purcell factor P12 [32]. As shown in Eq. (5), in the absence of the cavity, the microwave transition rates γi,j would be much smaller than exciton relaxation rates γi. The inner cavity damping rate is γc1=ωc12Q. Other damping rates γi,j are calculated from Eqs. (3)(5)), further modified by respective Purcell factors [17]. Due to the fact that the optical pump is a relatively short impulse which is spectrally wide, in contrast to [17] where continuous laser has pumped the system, it is possible to excite a set of upper levels n3±i, i=0,1,2 remaining in the adiabatic regime. Since for i0, the transitions n3±in2 are detuned from the cavity frequency ωc2, they are slower, resulting in significant populations N3±i. These populations reduce the overall efficiency of the system by occupying available space via the Rydberg blockade mechanism. This limitation is especially pronounced in the case of masers based on high excitonic levels n3, driven by very short pumping pulses. The pumping rate is given by Pr=Ppump/Eexciton(n3±i), where Eexciton is the energy needed to create the exciton and Ppump is the pump power. As mentioned in Sec. 2, the absorbed power is multiplied by the factor
b=1iVblockade(ni)Vtotal
which includes the effect of Rydberg blockade. According to the temperature dependence described by Eq. (4), for the ij transition with emission energy Δi,j, the emission rate scales with temperature as (1+ei,j/kT) due to the reduction of transition probability [36, 37]. This is especially pronounced in the case of higher excitonic states which are located closely together.

It should be stressed that in contrast to the continuous-wave system [17], the discussed system is characterized by a complex dynamics; the presence of a strong microwave field in the cavity causes a Stark shift, detuning the transitions from the respective cavity frequencies and changing the Purcell factors by up to 4 orders of magnitude and thus enabling or disabling the masing action. Moreover, the temperature of the system is not constant and has a nontrivial effect on the population dynamics. The optimal values of parameters such as geometric dimensions of the crystal, inner and outer cavity Q

factors, peak pump power and pumping pulse duration are highly dependent on the chosen excitonic levels.

Finally, we note that a masing action can be also obtained in a simpler, three level system which can be described by modifying Eq. (9), in particular omitting the N3 populations and instead applying pump to the ensemble of N2±i states. However, this approach allows for less flexibility regarding the choice of the excitonic states and utilizes one tunable cavity as opposed to two in the 4-level scheme.

 figure: Fig. 1

Fig. 1 Schematic depiction of the proposed system and excitonic energy level scheme.

Download Full Size | PDF

4. Three-level system

Let’s consider a 3 level system based on the modified setup shown on Fig. 1. The pump is tuned to the n2P state and masing occurs on the n2n1 transition. The system uses one microwave cavity tuned to the masing transition. By solving modified Eq. (9), one can obtain populations and emission power as a function of time.

 figure: Fig. 2

Fig. 2 Emission power in 3-level system, as a function of wavelength for T=10K and T=100K. The principal numbers n1, n2 are given in the brackets.

Download Full Size | PDF

Figure 2 shows the peak maser emission power as a function of wavelength for cavity Q=104, pump power P=10 W, pumping pulse duration τ = 20 ns and two selected temperatures. A significant number of state combinations (shown in brackets) exhibit an efficient microwave emission with maximum power of up to 10 mW. The wavelength depends mostly on the lower level n1 of the n2n1 transition. The systems where n2=n1+1 are the most efficient due to the high overlap of wavefunctions, resulting in a significant transition dipole moment (e.g., for the system of [1, 2] or [2, 3] states). However, in some cases the transition dipole moment is sufficiently large to cause emission before any significant population inversion can occur, so that not all possible combinations of states are present on Fig. 2. For the higher temperature T=100 K, the configurations with n1>5 become inaccessible due to a significant broadening of the spectral lines. Notably, in these conditions, the quantum number of the upper state n2 can reach values exceeding 10; however, in such a case, the pump laser excites not only the chosen state n2, but also a whole ensemble of nearby states, lowering the overall efficiency of the system. Nevertheless, there are quite a lot of accessible states that enable the maser action of various wavelengths.

 figure: Fig. 3

Fig. 3 (a) State populations and blockade volume occupied by excitons in maser system based on [3, 5] levels. (b) Emission power and transition detuning from the cavity caused by Stark shift. (c) Total absorbed and emitted energy.

Download Full Size | PDF

On Fig. 3(a) one can see the populations of the maser states for the states’ pair [3,5] at T=10 K and the same conditions as in Fig. 2 (e.g., τ = 20 ns, P = 10 W). The chosen 53 transition has a frequency ofω53=12.1 THz. The time t = 0 marks the maximum of the pump laser power. As the power increases, the upper state reaches a significant population and the condition of population inversion is established. The n2P exciton population increases until the fraction of crystal taken by Rydberg blockade reaches unity, e. g. the point where the medium is saturated and no more excitons can be created. This is possible because the cavity is initially detuned from the n2Pn1S transition, as shown on Fig. 3(b). This can be achieved by applying a constant electric field to the system, inducing the Stark shift. At t20 ns, the inversion is sufficiently high to start masing despite the detuning. As the microwave field builds up in the cavity, the Stark shift changes the n2Pn1S transition frequency ω21 so that it is closer to the cavity frequency ωc1. This, in turn, further enhances the emission. As a result, there is an exponential increase in the emitted field up to the moment when the inversion is destroyed. This mechanism, which can be seen as a form of passive Q-switching, is a key feature of our proposal; it is only possible due to the exceptionally high sensitivity of excitonic levels to the applied electric field. One can see on Fig. 3(a) that at t20 ns there is a sudden drop of the population N2 and corresponding emission power peak on Fig. 3(b). Notably, the maximum power is in the range of 200 mW, much higher than in the same system on Fig. 2. This means that the introduction of an initial detuning allows for a much higher population inversion. One can see on Fig. 3(b) that the considered system has a second, wider emission peak starting at t = 0. At this point, the population N1 is still relatively high, but significant pump power maintains a small population inversion, leading to a quasi - stationary regime where all populations remain relatively constant, up to the time t30 ns. The emission power is smaller than in the initial peak, but still considerable. Moreover, due to the much longer emission time, this peak represents the biggest contribution to the total emitted energy (Fig. 3(c)). Another notable feature is the second, narrow peak at t40 ns. This point corresponds to the condition where the masing transition is perfectly tuned to the cavity. The overall timescale of the whole processes (400 ns) is consistent with the experimental results with Rydberg atom masers [19]. The shortest emission peak has a duration of τ1 ns which corresponds to the spectral width of δω1 GHz. During the emission peaks, the field is E102 V/cm, which is significantly below the breakdown voltage of 105 V/cm. From Fig. 3(c) one can conclude that the energetic efficiency of the system is on the order of 104. This is caused mainly by the fact that single optical frequency photon creates up to one exciton, which in turn emits one microwave frequency photon. The total emitted energy Eemit10 GeV corresponds to 1.31012 microwave photons (E538 meV), which is comparable with the population of the upper state. Assuming that, apart from microwave emission, all the remaining energy is dissipated as heat, a single cycle increases the temperature of 1 mm 3 crystal by 101 K (the specific heat of 489 J kg 1 K 1 is assumed). In conclusion, even for high pumping intensity, the maximum power is not thermally limited.

 figure: Fig. 4

Fig. 4 (a) State populations and blockade volume occupied by excitons in maser system based on [7, 8] levels. (b) Emission power and transition detuning from the cavity caused by Stark shift. (c) Total absorbed and emitted energy.

Download Full Size | PDF

To get a better insight into the dynamics of the maser system based on the upper excitonic levels, let’s consider n1=7 and n2=8 states with a much lower emission frequency ω78=0.734 THz. To avoid excessive bleaching [1], the pump power is reduced to P=10 mW but the pulse duration is increased to τ = 80 ns. Interestingly, one can see on Fig. 4(a) that even in these conditions, the system reaches saturation with almost 40% volume taken by excitons. This is caused by the fact that n = 8 excitons are much larger than the earlier considered n = 5 ones. One can see that the maximum population is of the order of 1011. As mentioned earlier, the large transition dipole moment between upper excitonic states means that the emission is possible even in the condition of detuning from the cavity. As seen on Fig. 4 (b), ω87ωc1 at all times, which facilitates the build-up of a significant population inversion. In this system, the emission occurs at t50 ns, after the peak of the pump power. Again, one can see that the sudden increase of the output power is caused by a reduction of the detuning caused by the Stark shift. The maximum power is 0.2 mW, which is by 3 orders of magnitude lower than in the previous example, accordingly to the proportionally reduced pump power. In contrast to the previous system, the emission occurs after the pump has been switched off; as the population inversion is destroyed, the power decreases exponentially, which is marked on Fig 4 (b) by a straight line. The emission time τ10 ns is considerably longer than in the [3, 5] system, which yields a narrower emission spectrum with full width at half maximum δω100 MHz. Finally, despite the much lower pump power, the total absorbed and emitted energy (Fig. 4(c)) are only smaller by a factor of 250 due to a longer pump pulse duration. The overall efficiency is still in the range of 104. To sum up, the masing systems based on higher excitonic levels are characterized by slower dynamics and much lower peak emission power. On the other hand, masing in these setups is feasible even with a very low pump power.

5. Four-level system

A four level system consists of the upper, optically pumped n3P state, intermediate, metastable n2S state and lower n1P state, as shown on the Fig. 1. The setup consists of two microwave cavities tuned to n3n2 and n2n1 transitions. Separating the pumped and metastable state and introducing two cavities allow for more flexibility in tuning the system. Specifically, one can utilize high frequency, high power lower level masing transition n2n1 in a system which is highly susceptible to the electric field due to the strong Stark shift of upper n3 level altering its tuning to the outer cavity. This results in the maser’s operation in strong, passive Q-switching regime, leading to a very rich dynamics.

 figure: Fig. 5

Fig. 5 Emission power in 4-level system, as a function of wavelength for T = 10K and T = 100K. The principal numbers n1, n2, n3 are given in the brackets. For any set of n1 and n2, the combinations with lowest and highest n3 are shown.

Download Full Size | PDF

Figure 5 shows the maser emission power as a function of wavelength for a 4-level system with inner cavity Q1=104 and outer cavity Q2=105. The pump power is P=10 W and the pulse duration is again τ = 20 ns. One can see that the results are grouped in columns corresponding to the common values of n1, n2 resulting in the same emission frequency ω21. In comparison with Fig. 2, more systems based on upper states are present. This is caused by the introducing the third excitonic state and second cavity; the pumped level n3 remains relatively empty at all times due to the Purcell effect, which increases the probability of n3n2 transition. This means that every pump photon has a high probability of creating n3P exciton and the effect of Rydberg blockade is not very significant as long as n2 is relatively low. The important advantage over a 3-level system is that the middle state n2 is an S-exciton state, which is characterized by a relatively longer lifetime [26], helping to achieve significant population inversion n2n1. Additionally, the lower level n1P has a shorter lifetime than in 3-level system which is beneficial for keeping this level empty.

 figure: Fig. 6

Fig. 6 (a) State populations and blockade volume occupied by excitons in maser system based on [2–4] levels. (b) Emission power and transition detuning from the cavity caused by Stark shift. (c) Total absorbed and emitted energy.

Download Full Size | PDF

Figure 6 shows a typical dynamics of 4-level maser with small initial detuning of the inner cavity ωc1 and no detuning of the outer cavity ωc2. The population of the relatively stable intermediate state n2S reaches a maximum value of 1011, filling 1% of the available space (Fig. 6(a)). The emission starts early at t=18 ns, with maximum power of 100 mW (Fig. 6(b)). Both n1P and n3P states are affected by the Stark shift. The higher frequency ω21 is not as strongly affected as the ω32 due to the fact that the higher levels are more shifted. Additionally, the higher Q factor of the outer cavity means that its operating frequency range is narrower. This means that the dynamics of the system is mainly governed by the detuning of ω32 from the outer cavity frequency ωc2 (Fig. 6(b)). Specifically, after some time a significant microwave field builds up in the cavity and the following Stark shift prevents any further emission. As a result, masing occurs in the form of a series of bursts (τ1 ns). As mentioned above, this occurs for Rydberg blockade fraction as small as 0.01; this value can be increased by changing ωc2 and delaying the emission. However, the output power of the maser would still be limited by the Stark shift of n3 level.

One can conclude that the system dynamics is highly complex and strongly dependent on the initial detuning of the cavities. The oscillation of populations on Fig. 6(a) reflects the emission power spectrum. Apart from these fast changes, one can also see that the populations of n1P and n3P states decrease exponentially for t>0 while that of n2S remains almost constant due to its longer lifetime. Also, the fast decay of n1P population allows for higher pulse repetition rate than in three level systems. Finally, the energetic efficiency is once again on the order of 104 (Fig. 6(c)).

6. Conclusions

We have discussed and numerically analysed a proposal for a high power, pulsed maser based on Rydberg exciton states. Our results indicate that the masing action is feasible in many three- and four level systems under a wide range of conditions, even at temperatures of over 100 K and with relatively low Q-factor cavities. A generation of nanosecond pulses with peak power of over 200 mW and overall device efficiency of 104 is reported. Moreover, we show that the dynamics of the system is much richer than that for the continuous-wave maser [17]; due to the Stark shift of energy levels, the conditions for population inversion are highly sensitive to the microwave field present in the cavity, making the system highly complex. This self-detuning affects the dynamics in a manner similar to passive Q-switching, enabling or disabling the masing action depending on the power stored in the cavity. The impact of the Rydberg blockade and temperature on the maser performance is also discussed; three level systems based on low n excitons offer highest emission power while the four level systems based on upper states provide the greatest tuning flexibility and pulse repetition rate.

Funding

National Science Centre, Poland (project OPUS 2017/25/B/ST3/00817).

Acknowledgment

We wish to thank Roman Ciuryło for a valuable discussion and an enthusiastic encouragement.

References

1. T. Kazimierczuk, D. Fröhlich, S. Scheel, H. Stolz, and M. Bayer, "Giant Rydberg excitons in the copper oxide Cu 2O," Nature 514, 343 (2014). [CrossRef]   [PubMed]  

2. S. Zielińska-Raczyńska, G. Czajkowski, and D. Ziemkiewicz, "Optical properties of Rydberg excitons and polaritons," Phys. Rev. B 93, 075206 (2016). [CrossRef]  

3. S. Zielińska-Raczyńska, D. Ziemkiewicz, and G. Czajkowski, "Electro-optical properties of Rydberg excitons," Phys. Rev. B 94, 045205 (2016). [CrossRef]  

4. S. Zielińska-Raczyńska, D. Ziemkiewicz, and G. Czajkowski, "Magneto-optical properties of Rydberg excitons: Center-of-mass quantization approach," Phys. Rev. B 95, 075204 (2017). [CrossRef]  

5. J. Heckötter, M. Freitag, D. Fröhlich, M. Aßmann, M. Bayer, M. A. Semina, and M. M. Glazov, "Scaling laws of Rydberg excitons," Phys. Rev. B 96, 125142 (2017). [CrossRef]  

6. M. Aßmann, J. Thewes, D. Fröhlich, and M. Bayer, "Quantum chaos and breaking of all anti-unitary symmetries in Rydberg excitons," Nature Materials 15, 741–745 (2016). [CrossRef]  

7. V. Walther, R. Johne, and T. Pohl, "Giant optical nonlinearities from Rydberg excitons in semiconductor microcavities," Nat. Commun. 9, 1309 (2018). [CrossRef]   [PubMed]  

8. V. Walther, S. O. Kruger, S. Scheel, and T. Pohl, "Interactions between Rydberg excitons in Cu 2O," Phys. Rev. B 98, 165201 (2018). [CrossRef]  

9. M. Khazali, K. Heshami, and C. Simon, "Single-photon source based on Rydberg exciton blockade," J. Phys. B 50, 215301 (2017). [CrossRef]  

10. H. Kang, K. Wen, and Y. Zhu, "Normal or anomalous dispersion and gain in a resonant coherent medium," Phys. Rev. A 68, 063806 (2003). [CrossRef]  

11. Z. Zhang, J. Feng, X. Liu, J. Sheng, Y. Zhang, Y. Zhang, and M. Xiao, "Controllable photonic crystal with periodic Raman gain in a coherent atomic medium," Opt. Lett. 43, 919–921 (2018). [CrossRef]   [PubMed]  

12. D. Kleppner, H. C. Berg, S. B. Crampton, N. F. Ramsey, R. F. C. Vessot, H. E. Peters, and J. Vanier, "Hydrogen-Maser Principles and Techniques," Phys. Rev. 138, A972–A983 (1965). [CrossRef]  

13. A.E. Siegman, "Microwave Solid-State Masers" (McGraw-Hill,1964).

14. M. Oxborrow, J. D. Breeze, and N. M. Alford, "Room-temperature solid-state maser," Nature 488, 353–356 (2012). [CrossRef]   [PubMed]  

15. L. Jin, M. Pfender, N. Aslam, P. Neumann, S. Yang, J. Wrachtrup, and R.-B. Liu, "Proposal for a room-temperature diamond maser," Nat. Commun. 6, 8251 (2014). [CrossRef]  

16. J. D. Breeze, E. Salvadori, J. Sathian, NMcN. Alford, and C. W. M. Kay, "Continuous-wave room-temperature diamond maser," Nature 493, 25970 (2018).

17. D. Ziemkiewicz and S. Zielińska-Raczyńska, "Proposal of tunable Rydberg exciton maser," Opt. Lett. 43, 3742 (2018). [CrossRef]   [PubMed]  

18. D. Ziemkiewicz and S. Zielińska-Raczyńska, "Dynamically Steered Maser Action of Rydberg Excitons in Cu 2O," Phys. Status Solidi B, https://onlinelibrary.wiley.com/doi/abs/10.1002/pssb.201800503 (2019).

19. L. Moi, P. Goy, M. Gross, J. M. Raimond, C. Fabre, and S. Haroche, "Rydberg-atom masers. I. A theoretical and experimental study of super-radiant systems in the millimeter-wave domain," Phys. Rev. A 27, 4, 2043–2064 (1983). [CrossRef]  

20. M. Weissbluth, "Atoms and Molecules" (Academic Press, 1978).

21. T. Kitamura, M. Takahata, and N. Naka, "Quantum number dependence of the photoluminescence broadening of excitonic Rydberg states in cuprous oxide," J. Lumin. 192, 808–813 (2017). [CrossRef]  

22. H. Zhao, S. Wachter, and H. Kalt, "Effect of quantum confinement on exciton-phonon interactions," Phys. Rev. B 66, 085337 (2002). [CrossRef]  

23. S. Rudin, T. L. Reinecke, and B. Segall, "Temperature-dependent exciton linewidths in semiconductors," Phys. Rev. B 42, 11218 (1990). [CrossRef]  

24. H. Stolz, F. Schöne, and D. Semkat, "Interaction of Rydberg excitons in cuprous oxide with phonons and photons: optical linewidth and polariton effect," New J. Phys. 20, 023019 (2018). [CrossRef]  

25. T. Itoh and S. I. Narita, "Analysis of Wavelength Derivative Spectra of Exciton in Cu 2O," J. Phys. Soc. Jpn. 39, 140–147 (1975). [CrossRef]  

26. D. Fröhlich, A. Nöthe, and K. Reimann, "Observation of the Resonant Optical Stark Effect in a Semiconductor," Phys. Rev. Lett. 55, 1335 (1985). [CrossRef]   [PubMed]  

27. T. F. Gallagher, "Rydbeg atoms" (Cambridge University Press, 1994). [CrossRef]  

28. J. H. Hoogenraad and L. D. Noordam, "Rydberg atoms in far-infrared radiation fields. I. Dipole matrix elements of H, Li, and Rb," Phys. Rev. A 57, 4533 (1998). [CrossRef]  

29. J. Heckötter, M. Freitag, M. Fröhlich, M. Aßmann, M. Bayer, M. A. Semina, and M. M. Glazov, "High-resolution study of the yellow excitons in Cu 2O subject to an electric field," Phys. Rev. B 95, 035210 (2017). [CrossRef]  

30. I. Hiroyuki, H. Mizuhiko, U. Jun, M. Takao, T. Masahiro, T. Ken-ichiro, U. Masaro, and M. Kenjiro, "Hydrogen Maser," Journ. Natl. Inst. Inf. Commun. Technol. 50, 85 (2003).

31. D. M. Strayer, G. J. Dick, and J. E. Mercereau, "Performance of a superconducting cavity stabilized ruby maser oscillator," IEEE Trans. Magn. 23, 1624–1628 (1987). [CrossRef]  

32. E. M. Purcell, "Spontaneous emission probabilities at radio frequencies," Phys. Rev. 69, 674 (1946).

33. C. Sauvan, J. P. Hugonin, I. S. Maksymov, and P. Lalanne, "Theory of the Spontaneous Optical Emission of Nanosize Photonic and Plasmon Resonators," Phys. Rev. Lett. 110, 237401 (2013). [CrossRef]   [PubMed]  

34. K. F. Renk, "Basics of Laser Physics" (SpringerInternational Publishing, 2017).

35. J.C. Garrison and R.Y. Chiao, "Quantum optics" (Oxford University Press, 2012).

36. W. F. Krupke, M. D. Shinn, J. E. Marion, J. A. Caird, and S. E. Stokowski, "Spectroscopic, optical, and thermomechanical properties of neodymium- and chromium-doped gadolinium scandium gallium garnet," J. Opt. Soc. Am. B 3, 102–114 (1986). [CrossRef]  

37. S. E. Pourmand, N. Bidin, and H. Bakhtiar, "Effects of temperature and input energy on quasi-three-level emission cross section of Nd 3+:YAG pumped by flashlamp," Chin. Phys. B 21, 094214 (2012). [CrossRef]  

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 (6)

Fig. 1
Fig. 1 Schematic depiction of the proposed system and excitonic energy level scheme.
Fig. 2
Fig. 2 Emission power in 3-level system, as a function of wavelength for T=10K and T=100K. The principal numbers n1, n2 are given in the brackets.
Fig. 3
Fig. 3 (a) State populations and blockade volume occupied by excitons in maser system based on [3, 5] levels. (b) Emission power and transition detuning from the cavity caused by Stark shift. (c) Total absorbed and emitted energy.
Fig. 4
Fig. 4 (a) State populations and blockade volume occupied by excitons in maser system based on [7, 8] levels. (b) Emission power and transition detuning from the cavity caused by Stark shift. (c) Total absorbed and emitted energy.
Fig. 5
Fig. 5 Emission power in 4-level system, as a function of wavelength for T = 10K and T = 100K. The principal numbers n1, n2, n3 are given in the brackets. For any set of n1 and n2, the combinations with lowest and highest n3 are shown.
Fig. 6
Fig. 6 (a) State populations and blockade volume occupied by excitons in maser system based on [2–4] levels. (b) Emission power and transition detuning from the cavity caused by Stark shift. (c) Total absorbed and emitted energy.

Tables (1)

Tables Icon

Table 1. Transition dipole moments d n 1 n 2 between n 1 S (rows) and n 2 P (columns) states given in units of e a B.

Equations (9)

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

r n = 1 2 a B [ 3 n 2 l ( l + 1 ) ]
V b l o c k a d e = 3 10 16 n 7 [ mm 3 ] .
γ 0 P ( n ) = 24  meV 1 + 0.01 n 2 n 3
γ P ( n ) = γ 0 P ( n ) + γ A C T + γ L O [ exp  ( ω L O / k B T ) 1 ] 1 .
γ i j γ i ω i j 3 ω i 3 .
E = E g R y ( n δ p ) 2 ,
Δ E = 2.4 10 6 n [ 4.2 ( n 8.452 ) 2 + ( n 35.16 ) 3 ] F ( 1 + 2.2 / n ) [ eV ] ,
N 3 ± i t = b P r γ 3 ± i N 3 ± i γ 3 ± i , 2 N 3 ± i N 2 t = γ 3 ± i , 2 N 3 ± i i γ 2 N 2 γ 2 , 1 N 2 ( N 2 N 1 ) B W N 1 t = γ 2 , 1 N 2 + ( N 2 N 1 ) B W γ 1 N 1 W t = ω ( N 2 N 1 ) B W 2 γ c 1 W
b = 1 i V b l o c k a d e ( n i ) V t o t a l
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.