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

Measurement of deep-subwavelength emitter separation in a waveguide-QED system

Open Access Open Access

Abstract

In the waveguide quantum electrodynamics (QED) system, emitter separation plays an important role for its functionality. Here, we present a method to measure the deep-subwavelength emitter separation in a waveguide-QED system. In this method, we can also determine the number of emitters within one diffraction-limited spot. In addition, we also show that ultrasmall emitter separation change can be detected in this system which may then be used as a waveguide-QED-based sensor to measure tiny local temperature/strain variation.

© 2017 Optical Society of America

1. Introduction

Photon transport in a waveguide system coupled to quantum emitters, known as “waveguide-QED”, has attracted extensive interests because of its possible applications in quantum device and quantum information [1,2]. The static and dynamical solutions of the photon transport in the waveguide-QED system have been widely studied using various methods [3–26]. Many possible applications have also been proposed such as atomic mirror/cavity [27–30], single-photon frequency comb generation [31], single-photon diode [32–34], single-photon transistor [35–38], single-photon frequency converter [39–42], and quantum computation [43–45]. In addition, the waveguide-QED system is also a very good platform to study the many-body physics beacause long-range interaction is allowed in this system [46–48]. Aside from the usual dielectric waveguide [49] and photonic crystal waveguide [50], these theories can be also applied to study the propagation of surface plasmon along the nanowire [51] and the microwave photon along the superconducting transmission line [52–57].

In the waveguide-QED system, collective interactions betweeen the emitters are critical for its functionality and they largely depend on the emitter separation [8,22,31,58,59]. However, few has been discussed about how to measure the emitter separation in the waveguide-QED system especially when the emitters are in the deep-subwavelength scale. Besides, it is also an open question about how to determine the number of emitters coupled to the waveguide when the emitters are in a subwavelength region. Near-field scanning optical microscopy (NSOM) may be able to determine the number and positions of the emitters with high accuracy based on near-field point-by-point scanning [60]. However, the near-field technique is surface bound and has limited applications. If the emitters are embedded inside a photonic crystal waveguide with low loss to the free space, NSOM may fail to detect the signal.

In this paper, we show a method to measure the emitter separation embedded in or side-coupled to a 1D waveguide. Our method can measure the emitter separation even if they are in the deep-subwavelength scale. We also show that the number of emitters within one diffraction-limited spot can be detmined from the number of reflection peaks. Our method shown here may also be used as super-resolution biosensing [61]. In addition, since dipole-dipole energy shift is very sensitive to the emitter separation especially when they are in the deep subwavelength scale [62–65], we may also detect ultrasmall emitter separation change from the emission spectrum shift. This may then be used to probe tiny temperature or strain change with high sensitivity. This waveguide-QED-based temperature/strain sensor can also have very high spatial resolution because the sensing region has a deep-subwavelength size.

This paper is organized as follows: In Sec. 2 we show the schematic setup to measure the deep-subwavelength emitter separation and the emission spectra of the system. In Sec. 3, we show how to measure the deep-subwavelength emitter separation in the cases when γ = 0 and γ ≠ 0. In Sec. 4, we show how to measure the ultrasmall emitter separation change in this system which may then be used as a temperature sensor. In Sec. 5, we show how to measure the coupling strength for non-identical emitter case. Finally we summarize the results.

2. Model and spectrum

The schematic setup for measuring the deep-subwavelength emitter separation is shown in Fig. 1. Suppose that the two emitters have a spatial separation d. They can couple to each other via the guided and non-guided photon modes. The dipole-dipole coupling between the two emitters is distance-dependent and they can modify the scattering photon spectra of the waveguide-QED system. By monitoring the emitted photon spectrum, we can determine the emitter separation d even if d is much smaller than the resonant wavelength λ.

 figure: Fig. 1

Fig. 1 Measurement of deep-subwavelength emitter separation and separation change in a waveguide-QED system.

Download Full Size | PDF

The interaction Hamiltonian of this system in the rotating wave approximation is given by [31]

H=j=12k(gkjeikzjσj+akeiδωkjt+H.c.)+j=12qƛ(gqƛjeiqrjaqƛσj+eiδωqƛjt+H.c.).
In this Hamiltonian, the first term is the interaction between the emitters and the waveguide photon modes with coupling strength gkj=μjEk(rj)/ (μ̄j is the transition dipole moment of the jth emitter, E→k (r→j) is the waveguide mode amplitude at position r→j with zj being its z component along the waveguide direction, and ħ is the Planck constant), and the second term is the interaction between the emitters and the non-guided photon modes with coupling strength gqƛj=μjEqƛ(rj)/ (E→q→ƛ(r→j) is the nonguided mode amplitude at position r→j with wavevector q→ and polarization ƛ). δωkj=ωkωj is the detuning between the transition frequency ωj of the jth emitter and the frequency ωk of the guided photon with wavevector k, and δωqƛj=ωqƛωj is the detuning between the transition frequency of the jth emitter and the non-waveguide photon modes with polarization ƛ and wavevector qσj+=|ejg| is the emitter raising operator and ak(aqƛ) is the annihilation operator of a guided photon (non-guided photon).

When we send a single photon pulse into this system, the reflection and transmission amplitudes have been calculated in Ref. [31] and they are given by

r(δk)=e2ikz12M12(δk)eikdM11(δk)e2ikdM22(δk)M11(δk)M22(δk)M122(δk),
t(δk)=1M11(δk)+M22(δk)2M22(δk)cos(kd)M11(δk)M22(δk)M122(δk),
where k = ka + δk with ka = (ω1 + ω2)/2c, d = |r2r1| is the distance between two emitters, Mii(δk) = (1 + γii) + 2iωiδkvg)/Γi and M12(δk)=2V12eikd/Γ1Γ2. Here, Γi=2L|gkaj|/vg (γi=kj3μj2/3π0V) is the spontaneous decay rate of the ith emitter due to the guided (non-guided) photon modes where L is the quantization length of the waveguide and V is the quantization volumn of the non-guide modes. V12e ikad is the dipole-dipole interaction strength between the two emitters with V12 given by
V12=Γ1Γ22+3γ1γ22[ikad+1(kad)2+i(kad)3],
where the first term is the dipole-dipole interaction induced by the waveguide modes and the second term is the dipole-dipole interaction induced by the non-guided modes. Here, we assume that the incident photon has a transverse polarization (TE-mode) so that the induced transition dipole moment is perpendicular to the waveguide direction. From Eq. (4), we see that the dipole-dipole interaction is highly distance-dependent. In the following, we first show how to determine the deep-subwavelength emitter separation from the spectrum characteristics in the cases with or without non-guided photon modes. Then we show how to measure ultrasmall emitter separation change from the emission spectrum shift. Since we mainly consider the case when the emitters have deep-subwavelength distance, we can safely assume that the emitters feel the same local field and have the same decay rates, i.e, Γ1 = Γ2 ≡ Γ and γ1 = γ2γ. In Sec. V, we consider the case when Γ1 ≠ Γ2 and γ1γ2.

3. Measuring deep-subwavelength emitter separation

3.1. γ = 0

We first consider a high quality 1D waveguide where non-guided modes are negligible, i.e., γ = 0. In this case, M11(δk) = M22(δk) = 1 − 2iδkvg/Γ and M12 = eikd. The reflection amplitude is then given by

r(δk)=e2ikz12e2ikd(12iδkvg/Γ)(1+e2ikd)(2iδkvg/Γ)2e2ikd,
from which it is not difficult to see that the reflection vanishes under the condition
δωδkvg=Γ2tan(ka+δk)d.
Since δk is usually much less than ka, the reflection vanishes at the frequency approximately equals to −Γ tan(kad)/2 which is a function of emitter separation. The reflection dip frequency as a function of emitter separation is shown in Fig. 2(a) where the red asterisk symbol is the exact numerically calculated result from Eq. (6) and the black solid line is the fitting of −Γ tan(kad)/2. We can see that the theoretical prediction matches the result calculated by Eq. (6) very well. Hence, from the dip position Δdip we can determine the emitter separation d = /2 + λarctan−1(2Δdip/Γ)/2π where n = 0, 1, · · ·. The existence of n is because the dipole-dipole coupling induced by the guided modes is a periodical function. Additional procedure is required to fix n. For example, a gradient field can be applied to distinguish the results for different n.

 figure: Fig. 2

Fig. 2 (a) The reflection dip frequency as a function of emitter separation. The red asterisk symbol is the numerical calculated dip positions and the black solid line is the fitting of − tan(kad)/2. (b) The reflection spectra for three different subwavelength emitter separation (d = 0.05λ, 0.1λ, 0.15λ). (c) The reflection spectra for three emitter separations (d = 0.05λ, 0.55λ, 1.05λ) when a gradient electric field is applied. (d) The reflection spectra when there are 2, 3, and 4 emitters within one diffraction-limited spot. The emitter separation d = 0.05λ.

Download Full Size | PDF

The reflection spectra for three different emitter separations (d = 0.05λ, 0.1λ, 0.15λ) are shown in Fig. 2(b). Indeed, there is a reflection dip in all three reflection spectra and the dip positions are different for different emitter separations. When the emitter separation increases from d = 0.05λ to d = 0.15λ, the dip frequency is shifted from −0.162Γ to −0.688Γ. From the dip frequency −0.162Γ we can determine that the emitter separation is 0.05λ, and from −0.688Γ we can determine that the emitter separation is 0.15λ. Both results match the given value very well. However, we should mention that the solution is not unique because /2 + d also gives the same spectrum. This problem can be solved by applying a gradient electric or magnetic field to shift the transition frequency of the emitters. For a fixed gradient field, different emitter separation has different energy shift. For example, the reflection spectra when d = (n/2 + 0.05)λ with n = 0, 1, 2 are shown in Fig. 2(c) where the energy shift due to the gradient field is 2.0Γd/λ. When d = 0.05λ, the energy difference between the two emitters is 0.1Γ which is relatively small. The reflection spectrum is shown as the black solid curve in Fig. 2(c) where we see that the spectrum is similar to the case when there is no gradient field. However, when d = 0.55λ, the energy difference between the two emitters is 1.1Γ and the reflection spectrum (red dashed line) is quite different from the case without gradient field. There are two reflection peaks with similar shape and the separation between the two peaks is measured to be 1.1Γ. From this separation, we can determine the emitter separation d = 0.55λ. Similarly, when d = 1.05λ, there are also two reflection peaks with separation 2.1Γ from which we can determine that d = 1.05λ (blue dotted line).

When there are more than two emitters which are very close to each other, the spectrum can become more complicated and a full spectrum fitting may be required to extract all the emitter separations. Nonetheless, the number of emitters can be determined by simply counting the number of dips in the reflection spectra. The examples when there are 2, 3, and 4 emitters are shown in Fig. 2(d) where we can see that there is one dip for 2 emitters, two dips for 3 emitters, and three dips for 4 emitters. Thus the number of emitters is equal to the number of dips in the reflection spectra plus one. Alternatively, we can also count the number of peaks. Because a dip can be formed between two peaks, the number of peaks is the same as the number of emitters. The number of dips/peaks in the reflection spectrum can be interpreted as the degree of freedom of the system [66]. By adding one emitter into the system, we can add an energy sublevel into the one-photon excitation subspace and therefore add an additional interference channel to the system which will then add one dip or peak.

3.2. γ ≠ 0

We then consider the case when the non-guided modes are not negligible, i.e., γ ≠ 0. This can be applied to the case when the waveguide is lossy or the emitters are side-coupled to the waveguide. In this case, the reflection amplitude shown in Eq. (2) can be rewritten as

r(δk)=A(δk)M11(δk)+M12(δk)+B(δk)M11(δk)M12(δk)
where A(δk) = −e2ikz1 (1 + eikd)2/2 and B(δk) = −e2ikz1 (1 − eikd)2/2. For deep-subwavelength (dλ) region, δkd ≃ 0 and kdkad. We have A(δk) ≃ −e2ikz1 (1 + eikad)2/2, B(δk) ≃ −e2ikz1 (1−eikad)2/2, and M12(δk) ≃ 2V12e ikad/Γ. The denominators in Eq. (7) can be simplified as
M11(δk)±M12(δk)=(2/Γ){i[δkvgω±]+Γ±}
where ω± ≃ ±Im[V12eikad] is the collective frequency shift due to the dipole-dipole interaction and Γ± ≃ (Γ + γ)/2 ± Re[V12eikad] is the collective decay rate. Thus, there are two reflection peaks at the frequencies ω± with the full width at half maximum (FWHM) linewidth 2Γ±. In the deep-subwavelength region, the reflectivity at the superradiant peak ω+ is about Γ2/Γ+2 with Γ+ ≃ Γ + γ. From the reflection peak positions, the linewidths and the reflectivity at the superradiant peak, we can obtain the dipole-dipole energy shifts and the collective decay rates. Since both values depend on the emitter separation, it becomes possible that their separation can be extracted from the spectrum.

The dipole-dipole energy shift as a function of emitter separation is shown in Fig. 3(a) and 3(b), respectively. We can see that when the distance is much smaller than one wavelength, the energy shift is mainly due to dipole-dipole interaction induced by the non-guided modes (Im (V12nw)). However, when the distance is of the order of wavelength or larger, the energy shift is mainly induced by the waveguide modes (Im (V12w)) because the dipole-dipole energy shift due to the non-guided modes is vanishing. Therefore, it is possible to use Im (V12nw) to estimate the dipole-dipole energy shift when dλ. The collective decay rate as a function of emitter separation is shown in Fig. 3(b) where we can see that the superadiant and subradiant decay rate periodically change with the emitter separation. When the emitter separation is small which is the Dicke regime, the symmetric state has decay rate approaching Γ + γ while the antisymmetric state has decay rate approaching zero.

 figure: Fig. 3

Fig. 3 (a) The collective energy shifts as a function of emitter separation with Γ = 2γ. Solid red line: the total energy shift; Green dotted line: shift due to the non-waveguide photon modes; Blue dashed line: shift due to the waveguide photon modes. (b) The collective decay rates as a function of emitter separation. Thick solid black line: total decay rate of the |+〉 state; Thin solid red line: total decay rate of the |−〉 state; Green dotted line: decay due to the non-waveguide photon modes; Blue dashed line: decay due to the waveguide photon modes. (c) The reflection spectra for two different sub-wavelength emitter separations (d = 0.05λ: red solid line; d = 0.08λ: blue dashed line) when Γ = 2γ = 2γ0 (here γ0 is a reference decay rate). (d) The reflection spectra when there are 2, 3, and 4 emitters with neighboring emitter separation being 0.05λ and Γ = 5γ = 5γ0.

Download Full Size | PDF

The typical reflection spectrum is shown in Fig. 3(c) where we see that there are two reflection peaks with one superradiant peak and the other one subradiant peak. From the maximum reflectivity at the superradiant peak we can determine the ratio between Γ and γ. Together with the condition Γ+ + Γ = Γ + γ, we can then determine Γ and γ separately. The difference of the two reflection linewidths gives Γ+ − Γ = −2Re[V12eikad]. The reflection peak position ωp = Im[V12eikad]. From these two equations, we can determine the emitter separation by searching a separation d which minimizes the variance δS={[Re(V12eikad)(Γ+Γ)/2]2+[Im(V12eikad)ωp]2}/γ02.

In Fig. 3(c), we show the reflection spectra for two different emitter separations (d = 0.05λ and d = 0.08λ). For both spectra, the reflectivity at the superradiant peak is 0.44 from which we can obtain Γ2 /(Γ + γ)2 = 0.44. Thus, we can obtain Γ = 1.97γ which is very close to the given value 2γ. The spectrum when d = 0.05λ is shown as the red solid line. From the spectrum, we can measure the two peak positions to be ±23.388γ0 and their linewidths are 5.88γ0 and 0.12γ0, respectively. From the summation of the two linewidths, we have Γ + γ = 3γ0. Since Γ = 1.97γ, we can obtain Γ = 1.99γ0 and γ = 1.01γ0. From the difference of the two linewidths, we have Re [V12eikad] = 1.44γ0. From the peak position we have Im[V12eikad] = 23.388γ0. By searching the parameters d such that δS=(Re[V12eikad]1.44γ0)2+(Im[V12eikad]23.388γ0)2/γ02 is minimized, we obtain that d = 0.0502λ with variance δS = 0.003. We can see that the emitter separation is very close to the actual values 0.05λ with an error 0.4%.

The reflection spectra when d = 0.08λ is shown as the blue dashed line in Fig. 3(c). The superradiant peak is at frequency 5.732γ0 with the linewidth 5.76γ0. The subradiant peak is at frequency −5.752γ0 with the linewidth 0.32γ. From the summation of the two linewidths we have Γ + γ = 3.04γ0. We can then determine that Γ = 2.02γ0 and γ = 1.02γ0 which is also very close to the given values. From the difference of the two linewidths, we have Re [V12eikad] = 1.36γ0. The superadiant peak position is slightly different from the subradiant peak position. We can use their average to estimate the imaginary part of the dipole-dipole interaction, i.e., Im [V12eikad] = 5.742γ0. By least square fitting, we can determine that d = 0.0808λ with δS = 0.0003. The extracted emitter separation is very close to the real separation 0.08λ with an error being about 1%. Hence, the deep-subwavelength emitter separation can be extracted with very high accuracy.

Similar to the case of perfect waveguide, we can also determine the number of emitters within one diffraction-limited spot by simply counting the number of reflection peaks in the reflection spectra. The numerical examples for 2, 3, and 4 emitters are shown in Fig. 3(d) where we can see that the number of reflection peaks equals to the number of emitters. Here we count the number of peaks instead of dips is because the dips become very broad in this case and it is easier to count the peaks.

4. Measuring emitter separation change and temperature sensing

In the previous section, we show that deep sub-wavelength emitter separation in a waveguide-QED system can be measured from the emission spectrum. Here we show that we can also measure ultrasmall emitter separation change in a waveguide-QED system from the reflection spectrum shift.

From Eq. (4) we see that the dipole-dipole energy shift Ω12 ≈ 3γ/4(kad)3 when dλ. The frequency shift caused by emitter separation change is given by δω ≈ −3Ω12δd/d. When d is small, Ω12 can be very large and δω can be very large for even very small δd. In the fiber Bragg grating sensor, the frequency shift due to the effective waveguide expansion is given by δωFBG ≈ −ωaδd/λ [67]. The ratio of the sensitivity between our method and the FBG sensor is given by δω/δωFBG ∼ 3λΩ12/(ωad). To increase the sensitivity of our method, we need to reduce the emitter separation d and increase Ω12 as large as possible. Supposing that γ = 10−7ωa and d = 3 × 10−3λ, we have Ω12 ≈ 10−2ωa and the sensitivity is enhanced by about 10 times. For λ ∼ 1μm, d ∼ 3nm, so the sensing region has a nanometer size. Hence, our method shown here can have both high sensitivity and spatial resolution. Because the strain or temperature can modify the effective reflection index or cause thermal expansion of the waveguide, measurement of the effective emitter separation change may also be used to detect the tiny strain or temperature variation.

The refection spectrum shifts due to the emitter separation changes when d ≃ 0.01λ are shown in Fig. 4 where Fig. 4(a) is the superradiant reflection spectrum shift and Fig. 4(b) is the subradiant relfection spectrum shift. The black solid curves in Fig. 4(a) and 4(b) are the results when d = 0.01λ. When the emitter separation is reduced by an amount δd = 10−4λ, the center of the superradiant peak is shifted by 92γ (blue dashed line in Fig. 4(a)). From this frequency shift, we can determine that the effetive emitter separation change δd = δωd/2Ω12 ≃ 10−4λ. The effective length change caused by strain in a silica fiber for λ = 1.55μm is about 1.25pm/με, while it is 12.5pm/K for temperature change [67]. For λ = 1.550μm, δd = 155pm. which corresponds to a change 124 microstrain or 12.4K temperature. The minimum distance change we can measure from the superradiant peak shift is about 2 × 10−5λ (red dotted line in Fig. 4(a)) where the shift of the superradiant peak is about 18γ and the reflection spectrum has significant overlap with the original spectrum. In this case, it corresponds to a change of 25 microstrain or 2.5K temperature. However, if we probe the subradiant peak, the sensitivity can be much higher because the subradiant peak has much narrower linewidth. As shown in Fig. 4(b), when the emitter separation change δd ∼ 10−7λ, the subradiant peak shift is about 0.09γ. Despite that this frequency shift is small, it is well separated from the original reflection peak. Usually it is difficult to detect the subradiant peaks. However, if we use a laser with narrowband frequency to scan through the subradiant peaks, it can be more efficient. The minimum frequency shift we can distinguish is about 0.02γ which correspons to a separation change of 2 × 10−8λ. When λ = 1.55μm, it corresponds to a change of 0.03pm which corresponds to a change of 0.02 microstrain or 2mK temperature. In this example, if γ ∼ 100MHz such as in silicon vacancy center [68], the minimum frequency shift is about 2MHz. In practice, we can use a continuously tunable laser with narrow linewidth to probe this frequency shift [69]. However, in practice the atom in a crystal can have thermal vibration at finite temperature. The vibration amplitude at room temperature is of the order of 10−11 ∼ 10−12m [70]. Due to this vibration, the smallest temperature variance we can measure is limited to be about 0.1°C. At very low temperature, the phonons are mostly long-wavelength acoustic phonon where the nearby lattice site can have almost the same vibration direction and amplitude. In this case, the emitter separation can remain almost the same despite that they can vibrate a bit. Since the dipole-dipole interaction depends on their relative separation instead of their absolute positions, the effect of the thermal vibrations to the sensor can be reduced and the sensitivity can be increased further.

 figure: Fig. 4

Fig. 4 Measurement of ultrasmall emitter separation change from spectrum shift. (a) Superradiant peak shift. (b) Subradiant peak shift. The solid lines are the reflection spectra when d = 0.01λ. The red dotted and blue dashed lines are the reflection spectra for the separation changes shown in the figure. Here Γ = 10γ.

Download Full Size | PDF

Our waveguide-QED-based sensor may find important applications in various areas because temperature plays a crucial role in many areas such as material formation and biochemical reaction [71]. There are a number of techniques for measuring temperature. The conventional thermometers, such as thermography and thermocouples, suffer from a lack of sensitivity and spatial resolution. Fluorescent polymeric thermometer (FPT), which is based on the temperature-dependent fluorescence life time, can achieve high sensitivity [72]. However, FPT is working on the diffraction-limited scale and the spatial resolution is not very high. The fiber Bragg grating (FBG) sensor can also measure ultrasmall temperature change, but its spatial resolution is also diffraction-limited [67, 73–75]. In comparison, our proposed waveguide-QED-based temperature sensor has sensitivity the same as the FBG, but with much higher spatial resolution where the sensing region can be in the nanometer region. The temperature sensing based on nitrogen-vacancy (NV) center can have both high sensitivity (tens of milliKevin) and high spatial resolution (nanometer region) [76–79]. However, one disadvantage of NV-center-based temperature sensing is that the microwave and laser applied may significantly alter the original sample temperature. Although the sensitivity of our waveguide-QED-based temperature sensor may have less sensitivity than that based on NV-center, our method has its own advantage. The probe photon in our temperature sensor is at the single-photon level and it is mostly confined inside the waveguide. Only a small portion of the evanescent light can couple to the sensing region, so the modification of the sample temperature by the probe light is minimized.

5. Non-identical coupling strength

In the previous sections, we assumed that the emitters have the same coupling strength with both the waveguide and non-waveguide modes. This assumption is valid when the two emitters are spatially very close to each other, i.e., they can feel the same local field. However, when the two emitters are not in the sub-wavelength distance, the two emitters can have different coupling strengths with the guided and non-guided fields. In the section, we show a method to measure the coupling strengths and decay rates of the two emitters when they are different.

We have shown that the reflectivity of a single emitter at the resonant frequency is given by (1 + γ/Γ)−2, and the linewidth of reflection spectrum is given by Γ + γ [22]. From these two characteristic parameters, we can determine Γ and γ, respectively. For the two-emitter case, we can first apply a gradient electric or magnetic field to distinguish the two emitters. If the energy difference between the two emitters is greater than their coupling strength, they can be treated as separate emitters and then we can determine the two coupling strengths separately. For example, we assume that the distance between the two emitters is 2.1λ with Γ1 = Γ0, γ1 = 0.5Γ0, Γ2 = 1.5Γ, and γ2 = 0.9Γ0. By applying a gradient field 6Γ0/λ, the emission photon spectra are shown in Fig. 5 where the red solid line is the reflection spectrum and the blue dotted line is the transmission spectrum. There are two reflection peaks which correspond to the two emitters. The separation between the two peaks is 12.6Γ0 from which we can determine that the emitter separation is 2.1λ which matches the given value very well. The FWHM linewidth of the left reflection peak is 1.45Γ0 and the maximum reflectivity is 0.45. We therefore have Γ1 + γ1 = 1.45Γ0 and (1 + γ11)−2 = 0.45 from which we can determine that Γ1 = 0.97Γ0 and γ1 = 0.48Γ0 which also match the given values Γ0 and 0.5Γ0 respectively very well. Similarly, the FWHM linewidth of the right reflection peak is 2.46Γ0 and the maximum reflectivity is 0.41. We have Γ2 + γ2 = 2.46Γ0 and (1 + γ22)−2 = 0.41. We can then determine that Γ2 = 1.58Γ0 and γ2 = 0.88Γ0 which also match the given values very well.

 figure: Fig. 5

Fig. 5 The reflection and transmission spectra of two-emitter system under gradient field. The emitter separation is 2.1Γ0 and the coupling constants are given by Γ1 = Γ0, γ1 = 0.5Γ0, Γ2 = 1.5Γ, and γ2 = 0.9Γ0.

Download Full Size | PDF

6. Summary

In summary, we have proposed a method to determine the emitter separation in a waveguide-QED system even if the emitter separation is in the deep sub-wavelength scale. For a high quality photonic waveguide with negligible decay to the free space, the emitter separation can be deduced from the reflection dip position. For a waveguide with decay to the free space, the emitter separation can be determined from the dipole-dipole splitting. If there are more than two emitters which are very close to each other, we can also determine the number of emitters by simply counting the number of reflection dips or the reflection peaks. Moreover, we also show how to measure ultrasmall emitter separation change in the waveguide-QED system. This may then be used to measure the strain or temperature variation with both high sensitivity and spatial resolution. We also show how to measure the decay rates to the waveguide even if the emitters have different coupling constants. Our theory here may find important applications in designing the waveguide-QED-based device and sensor.

Funding

Qatar National Research Fund (QNRF) NPRP Grant No. 8-352-1-074; King Abdulaziz City for Science and Technology grant (KACST).

Acknowledgments

We thank X. Zeng, M. T. Cheng, and C. P. Sun for helpful discussion.

References and links

1. Z. Liao, X. Zeng, H. Nha, and M. S. Zubairy, “Photon transport in a one-dimensional nanophotonic waveguide QED system,” Phys. Scr. 91, 063004 (2016). [CrossRef]  

2. D. Roy, C. Wilson, and O. Firstenberg, “Colloquium: Strongly interacting photons in one-dimensional continuum,” Rev. Mod. Phys. 89, 021001 (2017). [CrossRef]  

3. J.-T. Shen and S. Fan, “Coherent photon transport from spontaneous emission in one-dimensional waveguides,” Opt. Lett. 30, 2001–2003 (2005). [CrossRef]   [PubMed]  

4. J.-T. Shen and S. Fan, “Coherent single photon transport in a one-dimensional waveguide coupled with superconducting quantum bits,” Phys. Rev. Lett. 95, 213001 (2005). [CrossRef]   [PubMed]  

5. V. Yudson and P. Reineker, “Multiphoton scattering in a one-dimensional waveguide with resonant atoms,” Phys. Rev. A 78, 052713 (2008). [CrossRef]  

6. T. Tsoi and C. Law, “Quantum interference effects of a single photon interacting with an atomic chain inside a one-dimensional waveguide,” Phys. Rev. A 78, 063832 (2008). [CrossRef]  

7. S. Derouault and M. Bouchene, “One-photon wave packet interacting with two separated atoms in a one-dimensional waveguide: Influence of virtual photons,” Phys. Rev. A 90, 023828 (2014). [CrossRef]  

8. D. Kornovan, A. Sheremet, and M. Petrov, “Collective polaritonic modes in an array of two-level quantum emitters coupled to an optical nanofiber,” Phys. Rev. B 94, 245416 (2016). [CrossRef]  

9. H. Zheng, D. J. Gauthier, and H. U. Baranger, “Waveguide qed: Many-body bound-state effects in coherent and fock-state scattering from a two-level system,” Phys. Rev. A 82, 063816 (2010). [CrossRef]  

10. D. Roy, “Two-photon scattering by a driven three-level emitter in a one-dimensional waveguide and electromagnetically induced transparency,” Phys. Rev. Lett. 106, 053601 (2011). [CrossRef]   [PubMed]  

11. C.-H. Yan, L.-F. Wei, W.-Z. Jia, and J.-T. Shen, “Controlling resonant photonic transport along optical waveguides by two-level atoms,” Phys. Rev. A 84, 045801 (2011). [CrossRef]  

12. Q. Li, L. Zhou, and C. P. Sun, “Waveguide quantum electrodynamics: controllable channel from quantum interference,” Phys. Rev. A 89, 063810 (2014). [CrossRef]  

13. S. Fan, Ş. E. Kocabaş, and J.-T. Shen, “Input-output formalism for few-photon transport in one-dimensional nanophotonic waveguides coupled to a qubit,” Phys. Rev. A 82, 063821 (2010). [CrossRef]  

14. K. Lalumiere, B. C. Sanders, A. F. van Loo, A. Fedorov, A. Wallraff, and A. Blais, “Input-output theory for waveguide qed with an ensemble of inhomogeneous atoms,” Phys. Rev. A 88, 043806 (2013). [CrossRef]  

15. G. Chen, Y.-C. Yu, X.-L. Zhuo, Y.-G. Huang, H. Jiang, J.-F. Liu, C.-J. Jin, and X.-H. Wang, “Ab initio determination of local coupling interaction in arbitrary nanostructures: Application to photonic crystal slabs and cavities,” Phys. Rev. B 87, 195138 (2013). [CrossRef]  

16. S. Xu and S. Fan, “Input-output formalism for few-photon transport: A systematic treatment beyond two photons,” Phys. Rev. A 91, 043845 (2015). [CrossRef]  

17. T. Shi and C. P. Sun, “Lehmann-symanzik-zimmermann reduction approach to multiphoton scattering in coupled-resonator arrays,” Phys. Rev. B 79, 205111 (2009). [CrossRef]  

18. M.-T. Cheng and Y.-Y. Song, “Fano resonance analysis in a pair of semiconductor quantum dots coupling to a metal nanowire,” Opt. Lett. 37, 978–980 (2012). [CrossRef]   [PubMed]  

19. M. Pletyukhov and V. Gritsev, “Scattering of massless particles in one-dimensional chiral channel,” New J. Phys. 14, 095028 (2012). [CrossRef]  

20. T. Shi, D. E. Chang, and J. I. Cirac, “Multiphoton-scattering theory and generalized master equations,” Phys. Rev. A 92, 053834 (2015). [CrossRef]  

21. Y. Chen, M. Wubs, J. Mørk, and A. F. Koenderink, “Coherent single-photon absorption by single emitters coupled to one-dimensional nanophotonic waveguides,” New J. Phys. 13, 103010 (2011). [CrossRef]  

22. Z. Liao, X. Zeng, S.-Y. Zhu, and M. S. Zubairy, “Single-photon transport through an atomic chain coupled to a one-dimensional nanophotonic waveguide,” Phys. Rev. A 92, 023806 (2015). [CrossRef]  

23. X. Zang and P. Lalanne, “Theoretical treatment of the interaction between two-level atoms and periodic waveguides,” Opt. Lett. 40, 3869–3872 (2015). [CrossRef]   [PubMed]  

24. A. Roulet and V. Scarani, “Solving the scattering of n photons on a two-level atom without computation,” New J. Phys. 18, 093035 (2016). [CrossRef]  

25. Z. Liao, H. Nha, and M. S. Zubairy, “Dynamical theory of single-photon transport in a one-dimensional waveguide coupled to identical and nonidentical emitters,” Phys. Rev. A 94, 053842 (2016). [CrossRef]  

26. Z. Chen, Y. Zhuo, and J. -T. Shen, “Exact dissipation model for arbitrary photonic Fock state transport in waveguide QED systems,” Opt. Lett. 42, 887 (2017). [CrossRef]   [PubMed]  

27. L. Zhou, H. Dong, Y.-x. Liu, C. P. Sun, and F. Nori, “Quantum supercavity with atomic mirrors,” Phys. Rev. A 78, 063827 (2008). [CrossRef]  

28. D. E. Chang, L. Jiang, A. Gorshkov, and H. Kimble, “Cavity qed with atomic mirrors,” New J. Phys. 14, 063003 (2012). [CrossRef]  

29. F. Fratini, E. Mascarenhas, L. Safari, J-Ph. Poizat, D. Valente, A. Aufféves, D. Gerace, and M. F. Santos, “Fabry-Perot Interferometer with Quantum Mirrors: Nonlinear Light Transport and Rectification,” Phys. Rev. Lett. 113, 243601 (2014). [CrossRef]   [PubMed]  

30. P.-O. Guimond, A. Roulet, H. N. Le, and V. Scarani, “Rabi oscillation in a quantum cavity: Markovian and non-markovian dynamics,” Phys. Rev. A 93, 023808 (2016). [CrossRef]  

31. Z. Liao, H. Nha, and M. S. Zubairy, “Single-photon frequency-comb generation in a one-dimensional waveguide coupled to two atomic arrays,” Phys. Rev. A 93, 033851 (2016). [CrossRef]  

32. V. M. Menon, W. Tong, F. Xia, C. Li, and S. R. Forrest, “Nonreciprocity of counterpropagating signals in a monolithically integrated sagnac interferometer,” Opt. Lett. 29, 513–515 (2004). [CrossRef]   [PubMed]  

33. Y. Shen, M. Bradford, and J.-T. Shen, “Single-photon diode by exploiting the photon polarization in a waveguide,” Phys. Rev. Lett. 107, 173902 (2011). [CrossRef]   [PubMed]  

34. D. Roy, “Few-photon optical diode,” Phys. Rev. B 81, 155117 (2010). [CrossRef]  

35. D. E. Chang, A. S. Sorensen, E. A. Demler, and M. D. Lukin, “A single-photon transistor using nanoscale surface plasmons,” Nat. Phys. 3, 807–812 (2007). [CrossRef]  

36. D. Witthaut and A. S. Sørensen, “Photon scattering by a three-level emitter in a one-dimensional waveguide,” New J. Phys. 12, 043052 (2010). [CrossRef]  

37. T. G. Tiecke, J. D. Thompson, N. P. de Leon, L. R. Liu, V. Vuletic, and M. D. Lukin, “Nanophotonic quantum phase switch with a single atom,” Nature 508, 241–244 (2014). [CrossRef]   [PubMed]  

38. O. Kyriienko and A. S. Sørensen, “Continuous-wave single-photon transistor based on a superconducting circuit,” Phys. Rev. Lett. 117, 140503 (2016). [CrossRef]   [PubMed]  

39. M. Bradford, K. C. Obi, and J.-T. Shen, “Efficient single-photon frequency conversion using a sagnac interferometer,” Phys. Rev. Lett. 108, 103902 (2012). [CrossRef]   [PubMed]  

40. M. Bradford and J.-T. Shen, “Single-photon frequency conversion by exploiting quantum interference,” Phys. Rev. A 85, 043814 (2012). [CrossRef]  

41. W.-B. Yan, J.-F. Huang, and H. Fan, “Tunable single-photon frequency conversion in a sagnac interferometer,” Sci. Rep. 3, 3555 (2013). [CrossRef]   [PubMed]  

42. Z. H. Wang, L. Zhou, Y. Li, and C. P. Sun, “Controllable single-photon frequency converter via a one-dimensional waveguide,” Phys. Rev. A 89, 053813 (2014). [CrossRef]  

43. F. Ciccarello, D. E. Browne, L. C. Kwek, H. Schomerus, M. Zarcone, and S. Bose, “Quasideterministic realization of a universal quantum gate in a single scattering process,” Phys. Rev. A 85, 050305 (2012). [CrossRef]  

44. H. Zheng and H. U. Baranger, “Persistent quantum beats and long-distance entanglement from waveguide-mediated interactions,” Phys. Rev. Lett. 110, 113601 (2013). [CrossRef]   [PubMed]  

45. V. Paulisch, H. J. Kimble, and A. González-Tudela, “Universal quantum computation in waveguide qed using decoherence free subspaces,” New J. Phys. 18, 043041 (2016). [CrossRef]  

46. J. S. Douglas, H. Habibian, C.-L. Hung, A. V. Gorshkov, H. J. Kimble, and D. E. Chang, “Quantum many-body models with cold atoms coupled to photonic crystals,” Nat. Photon. 9, 326–331 (2015). [CrossRef]  

47. H. Zheng, D. J. Gauthier, and H. U. Baranger, “Cavity-free photon blockade induced by many-body bound states,” Phys. Rev. Lett. 107, 223601 (2011). [CrossRef]   [PubMed]  

48. Y.-L. L. Fang and H. U. Baranger, “Waveguide qed: Power spectra and correlations of two photons scattered off multiple distant qubits and a mirror,” Phys. Rev. A 91, 053845 (2015). [CrossRef]  

49. B. Dayan, A. S. Parkins, T. Aoki, E. P. Ostby, K. J. Vahala, and H. J. Kimble, “A photon turnstile dynamically regulated by one atom,” Science 319, 1062–1065 (2008). [CrossRef]   [PubMed]  

50. D. Englund, A. Faraon, B. Zhang, Y. Yamamoto, and J. Vučković, “Generation and transfer of single photons on a photonic crystal chip,” Opt. Express 15, 5550–5558 (2007). [CrossRef]   [PubMed]  

51. A. V. Akimov, A. Mukherjee, C. L. Yu, D. E. Chang, A. S. Zibrov, P. R. Hemmer, H. Park, and M. D. Lukin, “Generation of single optical plasmons in metallic nanowires coupled to quantum dots,” Nature 450, 402–406 (2007). [CrossRef]   [PubMed]  

52. A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, R.- S. Huang, J. Majer, S. Kumar, S. M. Girvin, and R. J. Schoelkopf, “Strong coupling of a single photon to a superconducting qubit using circuit quantum electrodynamics,” Nature 431, 162 (2004). [CrossRef]   [PubMed]  

53. A. A. Abdumalikov, O. Astafiev, A. M. Zagoskin, Y. A. Pashkin, Y. Nakamura, and J. S. Tsai, “Electromagnetically induced transparency on a single artificial atom,” Phys. Rev. Lett. 104, 193601 (2010). [CrossRef]   [PubMed]  

54. I.-C. Hoi, C. M. Wilson, G. Johansson, T. Palomaki, B. Peropadre, and P. Delsing, “Demonstration of a single-photon router in the microwave regime,” Phys. Rev. Lett. 107, 073601 (2011). [CrossRef]   [PubMed]  

55. I.-C. Hoi, T. Palomaki, J. Lindkvist, G. Johansson, P. Delsing, and C. M. Wilson, “Generation of nonclassical microwave states using an artificial atom in 1d open space,” Phys. Rev. Lett. 108, 263601 (2012). [CrossRef]   [PubMed]  

56. A. F. van Loo, A. Fedorov, K. Lalumière, B. C. Sanders, A. Blais, and A. Wallraff, “Photon-mediated interactions between distant artificial atoms,” Science 342, 1494–1496 (2013). [CrossRef]   [PubMed]  

57. I.-C. Hoi, A. F. Kockum, T. Palomaki, T. M. Stace, B. Fan, L. Tornberg, S. R. Sathyamoorthy, G. Johansson, P. Delsing, and C. M. Wilson, “Giant Cross Kerr Effect for Propagating Microwaves Induced by an Artificial Atom,” Phys. Rev. Lett. 111, 053601 (2013). [CrossRef]  

58. X. Li and L. Wei, “Probing a single dipolar interaction between a pair of two-level quantum system by scatterings of single photons in an aside waveguide,” Opt. Commun. 366, 163– 167 (2016). [CrossRef]  

59. M.-T. Cheng, J. Xu, and G. S. Agarwal, “Waveguide transport mediated by strong coupling with atoms,” Phys. Rev. A 95, 053807 (2017). [CrossRef]  

60. E. Betzig and R. J. Chichester, “Single molecules observed by near-field scanning optical microscopy,” Science 262, 1422–1425 (1993). [CrossRef]   [PubMed]  

61. G.-Y. Chen, N. Lambert, Y.-A. Shih, M.-H. Liu, Y.-N. Chen, and F. Nori, “Plasmonic bio-sensing for the fenna-matthews-olson complex,” Sci. Rep. 7, 39720 (2017). [CrossRef]   [PubMed]  

62. Z. Ficek and S. Swain, Quantum interference and coherence: theory and experiments, vol. 100 (Springer Science & Business Media, 2005).

63. J.-T. Chang, J. Evers, M. O. Scully, and M. Suhail Zubairy, “Measurement of the separation between atoms beyond diffraction limit,” Phys. Rev. A 73, 031803 (2006). [CrossRef]  

64. Z. Liao, M. Al-Amri, and M. S. Zubairy, “Resonance-fluorescence-localization microscopy with subwavelength resolution,” Phys. Rev. A 85, 023810 (2012). [CrossRef]  

65. W. Feng, D.-W. Wang, H. Cai, S.-Y. Zhu, and M. O. Scully, “Electromagnetically induced transparency with superradiant and subradiant states,” Phys. Rev. A 95, 033845 (2017). [CrossRef]  

66. Y. Shen and J. -T. Shen, “Nanoparticle sensing using whispering-gallery-mode resonators: Plasmonic and Rayleigh scatterers,” Phys. Rev. A 85, 013801 (2012). [CrossRef]  

67. W. J. Bock, I. Gannot, and S. Tanev, Optical waveguide sensing and imaging (Springer Science & Business Media, 2007).

68. L. J. Rogers, K. D. Jahnke, T. Teraji, L. Marseglia, C. Muller, B. Naydenov, H. Schauffert, C. Kranz, J. Isoya, L. P. McGuinness, and F. Jelezko, “Multiple intrinsically identical single-photon emitters in the solid state,” Nat. Commun. 5, 4739 (2014). [CrossRef]   [PubMed]  

69. G. A. Ball and W. W. Morey, “Continuously tunable single-mode erbium fiber laser,” Opt. Lett. 17, 420–422 (1992). [CrossRef]   [PubMed]  

70. K. Lonsdale, “Vibration amplitudes of atoms in cubic crystals,” Acta Crystallogr. 1, 142–149 (1948). [CrossRef]  

71. B. B. Lowell and B. M. Spiegelman, “Towards a molecular understanding of adaptive thermogenesis,” Nature 404, 652–660 (2000). [PubMed]  

72. K. Okabe, N. Inada, C. Gota, Y. Harada, T. Funatsu, and S. Uchiyama, “Intracellular temperature mapping with a fluorescent polymeric thermometer and fluorescence lifetime imaging microscopy,” Nat. Commun. 3, 705 (2012). [CrossRef]   [PubMed]  

73. Y.-J. Rao, “In-fibre bragg grating sensors,” Meas. Sci. Technol. 8, 355 (1997). [CrossRef]  

74. A. D. Kersey, M. A. Davis, H. J. Patrick, M. LeBlanc, K. P. Koo, C. G. Askins, M. A. Putnam, and E. J. Friebele, “Fiber grating sensors,” J. Lightwave Technol. 15, 1442–1463 (1997). [CrossRef]  

75. B. Lee, “Review of the present status of optical fiber sensors,” Opt. Fiber Technol. 9, 57–79 (2003). [CrossRef]  

76. P. Neumann, I. Jakobi, F. Dolde, C. Burk, R. Reuter, G. Waldherr, J. Honert, T. Wolf, A. Brunner, J. H. Shim, D. Suter, H. Sumiya, J. Isoya, and J. Wrachtrup, “High-precision nanoscale temperature sensing using single defects in diamond,” Nano Lett. 13, 2738–2742 (2013). [CrossRef]   [PubMed]  

77. G. Kucsko, P. C. Maurer, N. Y. Yao, M. Kubo, H. J. Noh, P. K. Lo, H. Park, and M. D. Lukin, “Nanometre-scale thermometry in a living cell,” Nature 500, 54–58 (2013). [CrossRef]   [PubMed]  

78. H. Clevenson, M. E. Trusheim, C. Teale, T. Schroder, D. Braje, and D. Englund, “Broadband magnetometry and temperature sensing with a light-trapping diamond waveguide,” Nat. Phys. 11, 393–397 (2015). [CrossRef]  

79. I. Fedotov, S. Blakley, E. Serebryannikov, N. Safronov, V. Velichansky, M. Scully, and A. Zheltikov, “Fiber-based thermometry using optically detected magnetic resonance,” Appl. Phys. Lett. 105, 261109 (2014). [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 (5)

Fig. 1
Fig. 1 Measurement of deep-subwavelength emitter separation and separation change in a waveguide-QED system.
Fig. 2
Fig. 2 (a) The reflection dip frequency as a function of emitter separation. The red asterisk symbol is the numerical calculated dip positions and the black solid line is the fitting of − tan(kad)/2. (b) The reflection spectra for three different subwavelength emitter separation (d = 0.05λ, 0.1λ, 0.15λ). (c) The reflection spectra for three emitter separations (d = 0.05λ, 0.55λ, 1.05λ) when a gradient electric field is applied. (d) The reflection spectra when there are 2, 3, and 4 emitters within one diffraction-limited spot. The emitter separation d = 0.05λ.
Fig. 3
Fig. 3 (a) The collective energy shifts as a function of emitter separation with Γ = 2γ. Solid red line: the total energy shift; Green dotted line: shift due to the non-waveguide photon modes; Blue dashed line: shift due to the waveguide photon modes. (b) The collective decay rates as a function of emitter separation. Thick solid black line: total decay rate of the |+〉 state; Thin solid red line: total decay rate of the |−〉 state; Green dotted line: decay due to the non-waveguide photon modes; Blue dashed line: decay due to the waveguide photon modes. (c) The reflection spectra for two different sub-wavelength emitter separations (d = 0.05λ: red solid line; d = 0.08λ: blue dashed line) when Γ = 2γ = 2γ0 (here γ0 is a reference decay rate). (d) The reflection spectra when there are 2, 3, and 4 emitters with neighboring emitter separation being 0.05λ and Γ = 5γ = 5γ0.
Fig. 4
Fig. 4 Measurement of ultrasmall emitter separation change from spectrum shift. (a) Superradiant peak shift. (b) Subradiant peak shift. The solid lines are the reflection spectra when d = 0.01λ. The red dotted and blue dashed lines are the reflection spectra for the separation changes shown in the figure. Here Γ = 10γ.
Fig. 5
Fig. 5 The reflection and transmission spectra of two-emitter system under gradient field. The emitter separation is 2.1Γ0 and the coupling constants are given by Γ1 = Γ0, γ1 = 0.5Γ0, Γ2 = 1.5Γ, and γ2 = 0.9Γ0.

Equations (8)

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

H = j = 1 2 k ( g k j e i k z j σ j + a k e i δ ω k j t + H . c . ) + j = 1 2 q ƛ ( g q ƛ j e i q r j a q ƛ σ j + e i δ ω q ƛ j t + H . c . ) .
r ( δ k ) = e 2 i k z 1 2 M 12 ( δ k ) e i k d M 11 ( δ k ) e 2 i k d M 22 ( δ k ) M 11 ( δ k ) M 22 ( δ k ) M 12 2 ( δ k ) ,
t ( δ k ) = 1 M 11 ( δ k ) + M 22 ( δ k ) 2 M 22 ( δ k ) cos ( k d ) M 11 ( δ k ) M 22 ( δ k ) M 12 2 ( δ k ) ,
V 12 = Γ 1 Γ 2 2 + 3 γ 1 γ 2 2 [ i k a d + 1 ( k a d ) 2 + i ( k a d ) 3 ] ,
r ( δ k ) = e 2 i k z 1 2 e 2 i k d ( 1 2 i δ k v g / Γ ) ( 1 + e 2 i k d ) ( 2 i δ k v g / Γ ) 2 e 2 i k d ,
δ ω δ k v g = Γ 2 tan ( k a + δ k ) d .
r ( δ k ) = A ( δ k ) M 11 ( δ k ) + M 12 ( δ k ) + B ( δ k ) M 11 ( δ k ) M 12 ( δ k )
M 11 ( δ k ) ± M 12 ( δ k ) = ( 2 / Γ ) { i [ δ k v g ω ± ] + Γ ± }
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.