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

Broadband frequency conversion and “area law” in tapered waveguides

Open Access Open Access

Abstract

We propose to use a tapered waveguide to achieve optical frequency conversions on a photonic chip. We show that in tapered waveguides, the frequency conversion has a much broader bandwidth, which is proportional to the waveguide width difference as Δλ ∝ 1.7δw. More importantly, the conversion efficiency within the bandwidth is almost constant, which is favorable for ultrashort pulses. This simple but efficient design not only enables the conversion bandwidth to be engineered, but also is tolerant to fabrication error. We demonstrate an “area law” for the frequency conversion process that the integral of the conversion efficiency within the conversion bandwidth does not change against the waveguide lateral shape. This can be used as a general guideline for integrated nonlinear optical device designs. With our approach, high-efficiency and wavefront-keeping conversion for short pulses becomes possible on a photonic chip, which allows the applications for frequency comb and scalable on-chip information processing.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

In photonic integrated circuits (PICs), the nonlinear optical processes are greatly enhanced by the tightly confined optical fields [1–3]. These nonlinear optical processes have enabled various applications, such as frequency comb [4–6], supercontinuum generation [7,8], quantum photon sources [9, 10], and coherent frequency conversion [11, 12]. The coherent frequency conversion extends the operation bandwidth of information processing [13,14], and also enables the hybridization of different platforms [15]. For example, a lot of on-chip optical and electronic components work at visible and microwave wavelengths, while the communication between remote optical/electronic chips relies on fibers at near-IR wavelength. Via coherent frequency conversion, high-fidelity on-chip manipulation and high-efficiency communication can be realized simultaneously.

However, the frequency conversion in PICs faces experimental challenges due to its strict phase-matching condition. In a uniform waveguide, the frequency conversion efficiency strongly depends on wavelength due to material and mode dispersions. In other words, the frequency conversion bandwidth is limited, which prevents the applications of ultrashort pulses [16,17]. In practice, the efficient frequency conversion at target wavelength requires both coarse-tuning of the waveguide geometry in batches and fine-tuning of the dispersion via temperature, which highly increases the consumption and complexity of the experimental setup. Therefore, a design that can broaden the frequency conversion bandwidth and facilitate the experimental measurements is necessary.

It has been demonstrated that chirped periodic poling of nonlinear crystal [18] enables adiabatic frequency conversion and gives broadband conversion bandwidth. This configuration works well and is mostly adopted in free-space optics due to the bulky nonlinear crystal. While for PICs, implementing chirped periodic poling for on-chip waveguides is challenging. Instead, engineering the waveguide width like gratings is proposed [19–21], which is based on quasi-phase matching. These grating-like waveguides require only standard etching processes, and can efficiently but easily address those aforementioned challenges. On the other hand, a tapered waveguide whose profile is more simplified, is also capable of achieving broadband frequency conversion. The mechanism of broadband frequency conversion in these tapered waveguides is to tune the effective modal index continuously and to realize adiabatic mode conversion. It has been reported both theoretically in optical fiber [22] and experimentally in III–V waveguide [23]. However, the frequency conversion spectrum reported in Ref. [23] is still Lorentzian, rather than a flat-top shape as reported in the literature. Here, we provide a detailed guideline for the design of a tapered nonlinear waveguide, and focus more on the relation between conversion efficiency and bandwidth and the underlying physics. Real experimental validations for our proposal and Ref. [22] remain demanded.

In this article, we study the frequency conversion process between visible (600 nm) and telecom (1550 nm) wavelengths in a tapered waveguide. The frequency conversion spectrum is almost flat, whose bandwidth is proportional to the waveguide width difference between input and output ports. We refer the conversion efficiency integral over the entire spectrum as “area.” In the low conversion efficiency regime (e.g., weak pump or short waveguide length), the “area” for tapered waveguide is the same as that for uniform waveguide, indicating the broad bandwidth is at the expense of peak efficiency. While in the high conversion efficiency regime (e.g., strong pump or long waveguide length), the tapered waveguide outperforms the uniform waveguide in both bandwidth and efficiency. This is because the peak conversion efficiency of uniform waveguide easily saturates. Such tapered waveguide is less prone to fabrication imperfections, which makes it practical for experimental implementation. In particular, it can be used for the frequency conversion for ultrashort pulses.

2. Schematic and theory

We investigated the sum frequency generation (SFG) process in tapered aluminum nitride (AlN) waveguide. AlN is an excellent candidate for CMOS-compatible quantum PICs, with transparency across the ultraviolet to infrared spectrum and strong second-order nonlinearity of χ(2) = 4.7 pm/V [2,24,25]. In this work, the signal, pump, and idler wavelengths are chosen as λ1 = 1550 nm, λ2 = 980 nm, and λ3 = 600.4 nm, respectively. This SFG process bridges the visible and telecom wavelengths, and benefits the interconnection between different optical platforms.

Figure 1(a) gives the schematic illustration of a tapered AlN waveguide sitting on SiO2 buffer, with waveguide height h and waveguide length L. The waveguide width varies linearly along z-axis as w(z)=w0+δwL(zL2), where w(L2)=w0 is the central width and δw is the width difference between input and output ports. The waveguide eigenmodes were calculated with finite element method (COMSOL Multiphysics), as displayed in the insets of Fig. 1(a). Herein, the signal and pump photons are in 1st-order transverse-electric (TE) modes, and the idler photon is in 3rd-order TE mode. Note that, for idler photon, the 1st- and 2nd-order TE modes exist as well, but they were not included in our study owing to large phase mismatching and parity conservation. As shown in Fig. 1(b), we studied the modal effective refractive indices neff of the three eigenmodes as a function of waveguide width w. At w0 = 0.773 μm (denoted with gray dashed line), the phase mismatching between signal, idler, and pump photons is Δβ = n1k1 + n2k2n3k3 = 0. Here, nm (m = 1, 2, 3) and km are the modal effective index and free-space wave vector for signal, idler, and pump photons, respectively. In the following studies, we will carry out analytic calculations based on the effective refractive indices neff(w) fitted from Fig. 1(b).

 figure: Fig. 1

Fig. 1 (a) Schematic illustration of the tapered waveguide deposited on silica buffer whose width varies along the light propagation direction as w(z), with h = 300 nm. Insets: typical eigenmode distributions for signal, pump, and idler photons. (b) The effective refractive indices neff of eigenmodes at signal (green), pump (red), and idler (blue) wavelengths as a function of waveguide width w. Gray dashed line indicates the point where the phase matching condition is satisfied.

Download Full Size | PDF

Under the slowly varying amplitude approximation [26] and undepleted pump condition, the Hamiltonian of such coherent two-mode conversion in rotating coordinates can be represented as:

=(Δβ2g*gΔβ2).
Here, the coupling strength g is proportional to the mode overlap between signal, pump, and idler photons, and g2 is proportional to the average pump power P2. In the tapered waveguide, both Δβ and g vary along the z-axis. On the other hand, the evolution of Am (m = 1, 2, 3) which is the square root of photon number follows Schrödinger equation as:
iddz|Φ(z)=(z)|Φ(z),
with |Φ(z)〉 = {A1(z), A3(z)}T. Therefore, the effect of (z) on this system from 0 to z can be described with a transfer matrix as follows:
|Φ(z)=|Φ(0)=𝒯ei0z(s)ds|Φ(0).
Here, 𝒯 is the position-ordering Dyson operator. Assuming the initial state is |Φ(0)〉 = {1, 0}T, A3 in the final state becomes A3(z) = 21, where 21 is the off-diagonal matrix element of . Then the nonlinear conversion efficiency is obtained as η=|A3(z)A1(0)|2=|21|2. Due to the conservation of photon numbers (|Am|2) during the frequency conversion, we have maximum efficiency of η = 1. Here, the transfer matrix =𝒯ei0z(s)ds is numerically calculated with the fourth-order Runge-Kutta method.

Choosing w0 = 0.773 μm, as the waveguide width increases from w <w0 to w> w0, the phase mismatching is tuned from Δβ > 0 to Δβ < 0, which results in the broadband frequency conversion presented as follows [27].

3. Broadband frequency conversion

We calculated the conversion spectra of the SFG process in tapered AlN waveguide, as shown in Fig. 2(a). We also included the uniform waveguide (δw = 0) case for comparison. With fixed pump wavelength, the central wavelengths of the spectra do not change against δw. But the conversion bandwidth, which is defined as the full width at half maximum of the conversion spectrum, is much broader for the tapered waveguide. This will definitely benefit the frequency conversion of pulsed lasers and the applications in frequency division multiplexing. Besides, the conversion efficiency η within the bandwidth is almost flat, which is detailed explained in Appendix A. Thus, the waveform of a pulsed laser can be maintained due to the efficient conversion of all frequency components. Note that, we used P2 ∼ 1 W in our calculations, which is a realistic value for a short pulsed pump. We also provided the phase of the complex conversion coefficient in Appendix B, which is not flat but can be compensated via linear dispersive devices [28].

 figure: Fig. 2

Fig. 2 (a) Nonlinear conversion spectra of λ3 for δw = 0, 4, 8 nm, L = 10000 μm, and P2 = 1 W. (b) Bandwidth Δλ extracted from conversion spectra as a function of δw.

Download Full Size | PDF

We then extracted the bandwidth Δλ from the conversion spectra, and plotted it versus the waveguide width difference δw in Fig. 2(b) (red dots). As we can see, Δλ increases linearly against δw. We also reproduced this linear dependence in analytic manner, which is denoted as blue line in Fig. 2(b) and shows good agreement with the numerical data. Now we explain the linear dependence analytically as follows.

The conversion bandwidth of nonlinear optical processes can be calculated as:

Δλδwd(Δβ)dw/d(Δβ)dλ,
where d(Δβ)dw and d(Δβ)dλ are mode and material dispersions, respectively. For a tapered waveguide, the phase mismatching difference between input and output ports δβ) is approximately proportional to the waveguide width difference δw. And differentiating Δβ over the target wavelength (λ3), we have d(Δβ)dλ=2π(n3n1)/λ32. Substituting mode and material dispersions into Eq. (4), we are able to obtain the linear dependence of Δλ on δw. Details about derivations are provided in Appendix C.

Despite the broadband frequency conversion, the tapered waveguide also enables better tolerance to fabrication errors. A detailed study about this can be found in Appendix D.

4. Frequency conversion saturation

According to Fig. 2(a), the broad conversion bandwidth in tapered waveguide is at the expense of peak conversion efficiency ηpeak, which refers to the conversion efficiency at central wavelength of the spectra. Now we show that this limitation is only applicable to the low conversion efficiency regime (e.g., short interaction distance or weak pump).

We first integrated η over the entire spectrum for L = 1000 μm [cyan columns in Fig. 3(a)]. The integral ∫ ηdλ is the same for uniform (δw = 0) and tapered (δw > 0) waveguides, and can be explained analytically as follows. For uniform waveguide, η=|g|2L2sinc2(LΔβ2/4+|g|2) [26]. Then, the integral in the weak coupling regime (gL ≪ 1) can be expressed as:

ηdλ2π|g|2L/d(Δβ)dλ.
For tapered waveguide, according to the Landau-Zener transition [27], the transition probability (i.e., the frequency conversion efficiency) can be written as [18]:
η=1e2π|g|2|dΔβ/dz|2π|g|2|dΔβ/dz|,
under the condition |g|2 ≪ |dΔβ/dz|. Here the conversion efficiency is an approximation, because g slightly changes along z-axis for tapered waveguide. We then also obtain the integral asymptotically as:
ηdλ2π|g|2d(Δβ)dλdz=2π|g|2L/d(Δβ)dλ,
which is the same as Eq. (5). The analytical result of ∫ ηdλ = 0.1114 for L = 1000 μm well demonstrates consistence with the numerical data. Detail about this derivation is provided in Appendix C.

 figure: Fig. 3

Fig. 3 (a) Integral of η over the entire spectrum as a function of δw for L = 1000 μm, L = 10000 μm, and P2 = 1 W. (b) Integral conversion efficiency as a function of P2 for δw = 0 and δw = 4 nm with L = 1000 μm. Light blue rectangle highlights the parameter range where the “area law” holds (P2 < 6 W).

Download Full Size | PDF

We then did integration for L = 10000 μm [blue columns in Fig. 3(a)]. Obviously, ∫ ηdλ for tapered waveguide is much larger than that for uniform waveguide, but still obeys Eq. (7). The discrepancy for uniform waveguide originates directly from the saturation of ηpeak. We will interpret this later by combining with the effect of pump power P2.

In Fig. 3(b), ∫ ηdλ as a function of P2 is plotted for uniform and tapered waveguides. Under weak pump condition, ηpeak are small for both cases. Thus, Eqs. (5) and (7) hold perfectly within the light blue rectangular region. Note here, ∫ ηdλ linearly depends on P2 since |g|2P2. As the pump power increases (P2 > 6 W), the linear dependence breaks down for uniform waveguide. Because ηpeak which contributes most to ∫ ηdλ becomes saturated. In contrast, for tapered waveguide, smaller ηpeak results in higher saturation threshold P2th for pump power. Accordingly, the tapered waveguide also has higher saturation threshold Lth for interaction distance. Therefore, the tapered waveguide has better performance than uniform waveguide in the high conversion efficiency regime (e.g., long interaction distance or strong pump). More details can be found in Appendix E.

5. Broadband excitation

As having been demonstrated in Fig. 2, broadband idler photons can be generated in the tapered waveguide when the input is fixed. Inspired by this, we now think about whether broadband input can be converted in the tapered waveguide. So we record the conversion efficiency of idler photons when both signal and pump wavelengths are swept. As shown in Fig. 4(a), in the uniform waveguide, the efficient frequency conversion happens only when the phase matching condition is fulfilled. Except for the narrow conversion peak, the shape of the conversion spectrum is also sinc-function, similar to the spectrum shown in Fig. 2(a). Furthermore, the conversion spectrum in the tapered waveguide is shown in Fig. 4(b). As can be seen, the spectrum bandwidth is much broader, and the spectrum shape is also consistent with those in Fig. 2(a). We should expect that, as the waveguide width difference δw increases, the operational bandwidth of signal and pump wavelengths can be further broadened. The efficient conversion of broadband inputs demonstrates that the tapered waveguide is very suitable for applications related to pulsed lasers. For example, we can improve the frequency conversion efficiency with pulsed pump, which has higher peak power compared to continuous-wave laser with the same average power. On the other hand, we can also achieve the efficient frequency conversion of pulsed signal while maintaining its pulse shape, because of the almost constant conversion efficiency across the frequency spectrum.

 figure: Fig. 4

Fig. 4 (a) Conversion efficiency η of idler photons as a function of signal λ1 (x-axis) and pump λ2 (y-axis) wavelengths, with L = 10000 μm and δw = 0 nm. (b) Conversion efficiency η of idler photons as a function of signal λ1 (x-axis) and pump λ2 (y-axis) wavelengths, with L = 10000 μm and δw = 4 nm.

Download Full Size | PDF

6. Conclusion

To summarize, we report the broadband frequency conversion process in tapered AlN waveguide. By controlling the waveguide width, we achieve near-constant frequency conversion efficiency within a broad bandwidth (100 nm is achievable). In the low conversion efficiency regime (e.g., low pump power and short interaction distance), the integral of the efficiency over the entire bandwidth ∫ ηdλ linearly depends on the waveguide length and pump power, and is the same as that for uniform waveguide. It describes the trade-off between the peak conversion efficiency and conversion bandwidth, and implies that the bandwidth broadening in tapered waveguide is at the expense of the peak conversion efficiency. However, the linear dependence of ∫ ηdλ on the waveguide length and pump power quickly breaks down for uniform waveguide in the high conversion efficiency regime (e.g., high pump power and long interaction distance), due to the saturation of the peak conversion efficiency. While for tapered waveguide, this linear dependence is valid within a larger parameter range. Therefore, the tapered waveguide is more suitable for the frequency conversion of pulses, which have high peak power and broadband spectra. For example, we can use pulses to achieve both relatively high frequency conversion efficiency and broad bandwidth.

Our tapered structure enables broad conversion bandwidth, relaxes phase matching requirements, and is robust against the experimental imperfections in fabrication or input wavelengths. Such compact structure will find applications in the scalable on-chip information processing technology, especially for ultra-fast pulse optics and frequency conversion of combs. Besides, the design can be generalized to other materials, such as gallium arsenide, gallium nitride, and lithium niobate [29].

Appendices

A. Mechanism for broadband conversion

The underlying mechanism of flat and broad conversion bandwidth shown in Fig. 2(a) can be explained as follows. Firstly, we consider a waveguide segment with length δz (δzL), and the pump light is fixed with wavelength λ2.

At the starting position z of the waveguide segment it has width of w(z), where idler photons λ3 are phase matched with signal photons λ1. And their frequency conversion efficiency is η.

As the photons propagate, at position z′ = z + δz, the waveguide width becomes w(z′). And the phase matching condition is fulfilled for photons with wavelengths of λ′1 and λ′3. Their frequency conversion efficiency is η′. Note that, since the photons λ1 and λ3 are no longer phase matched at z′, the idler photons λ3 cannot be converted back to λ1, thus maintaining the efficiency η and exiting the waveguide together with photons λ′3.

Assuming that the waveguide segment is uniform within its length, i.e., w(z) = w(z′), the frequency conversion efficiency within the length δz under phase matched condition is [26]:

η=|g|2δz2sinc2(|g|δz),
where g is the nonlinear coupling strength as defined in the main text. Since g at position z for photons λ3 is almost equal to g′ at position z′ for photons λ′3, we have η = η′. Therefore, the nonlinear conversion spectrum in a tapered waveguide is flat within its broad bandwidth.

B. Phase of the complex conversion coefficient

In Fig. 2(a), the conversion efficiency (or the amplitude of complex conversion coefficient) is shown. In fact, to preserve the shape of original pulse, the conversion coefficient should also be flat in phase over the whole spectrum. Here in Fig. 5, we plot the phase spectrum of the complex conversion coefficient for different waveguide width difference δw. It shows that the phase is not constant within the conversion bandwidth. But a larger δw indeed helps to increase the bandwidth of phase. Such phase delay can be easily compensated via linear dispersive devices on a photonic chip [28].

 figure: Fig. 5

Fig. 5 The phase spectrum of the complex conversion coefficient for different waveguide width difference δw, with L = 1000 μm and P2 = 1 W.

Download Full Size | PDF

C. Derivations of equations

In this section, we provide the derivations of Eqs. (4) and (7) in more details.

For Eq. (4), we start from:

dλdw=Δλδw.
Then we obtain the bandwidth Δλ as:
Δλ=δwdλdw=δω1/dw1/dλ=δωd(Δβ)dw/d(Δβ)dλ.
Here, the term d(Δβ)/dλ is the differentiation of the phase mismatch over the target wavelength λ3. It is derived as follows.

From the definition of phase mismatch Δβ = n1k1 + n2k2n3k3, we have Δβ as a function of λ1, λ2, and λ3:

Δβ(λ1,λ2,λ3)=2π(n1λ1+n2λ2n3λ3).
According to the energy conservation principle:
2πcλ1+2πcλ2=2πcλ3,
we replace λ1 with λ2 and λ3 as:
Δβ(λ2,λ3)=2π(n2n1λ2n3n1λ3).
Assuming that the pump wavelength is fixed, we obtain the differentiation of Δβ over λ3 as:
d(Δβ)dλ3=2π(n3n1)λ32.
For Eq. (7), with Eq. (6), we derive it as:
ηdλ2π|g|2|d(Δβ)/dz|dλ=2π|g|2|d(Δβ)/dλ|dz=2π|g|2L/d(Δβ)dλ,
where the integral range starts from z = 0 to z = L.

D. Impact of fabrication errors

According to the practical experiences of experiments on photonic integrated circuits, a waveguide with slowly varying width can be fabricated, but there might be slight deviation of the absolute waveguide width. In this section, we present the effect of fabrication errors on the frequency conversion process, in which the waveguide width is modeled as:

w(z)=(w0+Δw)+δwL(zL2),
with Δw being the deviation of waveguide width due to imperfect fabrication.

We first study the effect of Δw on the conversion spectrum of idler photons. As shown in Fig. 6(a), Δw shifts the entire conversion window to longer/shorter wavelength, depending on it is positive or negative. Despite the wavelength center, the shape and bandwidth of the spectrum are not changed. It indicates that, the idler photons can still be efficiently converted to when the waveguide width is shifted, thanks to the broad conversion bandwidth.

 figure: Fig. 6

Fig. 6 (a) The frequency conversion spectrum of idler photons for different waveguide width deviation Δw, with δw = 4nm, L = 10000μm, and P2 = 1W. (b) The frequency conversion efficiency at 600.4 nm as a function of Δw for different waveguide width difference δw, with L = 10000μm and P2 = 1W.

Download Full Size | PDF

Then we study the tolerance of conversion efficiency to Δw for waveguides with different δw. In this case, we fix the wavelength of idler photons at 600.4 nm. As shown in Fig. 6(b), as the width difference δw increases, the tolerance to Δw also increases, and is approximately equal to ±δw/2. Meanwhile, the average conversion efficiency drops because of a larger δw.

These results validate our claim that the tapered waveguide is beneficial for imperfect fabrication, as a consequence of the broadband frequency conversion.

E. Saturation of the frequency conversion

At the core of the good performance demonstrated in Fig. 3 is the much lower saturation threshold in tapered waveguide. We elaborate this by comparing the frequency conversion spectra for uniform (δw = 0) and tapered (δw = 4nm) waveguides under different pump power. As shown in Fig. 7, when the pump power is increased by 5-fold, the flat-top spectrum of tapered waveguide has an average conversion efficiency increased by 3.9-fold. While the uniform waveguide has a sinc-function spectrum, with the peak conversion efficiency increased by 1.5-fold only. It unambiguously shows that the uniform waveguide is more easily saturated.

 figure: Fig. 7

Fig. 7 The frequency conversion efficiency under pump power of P2 = 1W and P2 = 5W, for uniform waveguide (a) and tapered waveguide (b), respectively.

Download Full Size | PDF

Funding

National Key Research and Development Program (2016YFA0301300); National Natural Science Foundation of China (NSFC) (61590932, 11774333, 61505195, 11874342); Anhui Initiative in Quantum Information Technologies (AHY130300); Strategic Priority Research Program of the Chinese Academy of Sciences (XDB24030601); Fundamental Research Funds for the Central Universities.

Acknowledgments

This work is partially carried out at the USTC Center for Micro and Nanoscale Research and Fabrication. Hong X. Tang acknowledges supports from the David & Lucile Packard Fellowship in Science and Engineering.

References

1. D. J. Moss, R. Morandotti, A. L. Gaeta, and M. Lipson, “New cmos-compatible platforms based on silicon nitride and hydex for nonlinear optics,” Nat. Photonics 7, 597 (2013). [CrossRef]  

2. C. Xiong, W. H. Pernice, X. K. Sun, C. Schuck, K. Y. Fong, and H. X. Tang, “Aluminum nitride as a new material for chip-scale optomechanics and nonlinear optics,” New J. Phys. 14, 095014 (2012). [CrossRef]  

3. L. Caspani, C. Xiong, B. J. Eggleton, D. Bajoni, M. Liscidini, M. Galli, R. Morandotti, and D. J. Moss, “Integrated sources of photon quantum states based on nonlinear optics,” Light. Sci. Appl. 6, e17100 (2017). [CrossRef]  

4. V. Brasch, M. Geiselmann, T. Herr, G. Lihachev, M. H. Pfeiffer, M. L. Gorodetsky, and T. J. Kippenberg, “Photonic chip–based optical frequency comb using soliton cherenkov radiation,” Science 351, 357–360 (2016). [CrossRef]   [PubMed]  

5. A. Ganesan, C. Do, and A. Seshia, “Phononic frequency comb via intrinsic three-wave mixing,” Phys. Rev. Lett. 118, 033903 (2017). [CrossRef]   [PubMed]  

6. X. X. Xue, F. Leo, Y. Xuan, J. A. Jaramillo-Villegas, P.-H. Wang, D. E. Leaird, M. Erkintalo, M. H. Qi, and A. M. Weiner, “Second-harmonic-assisted four-wave mixing in chip-based microresonator frequency comb generation,” Light. Sci. Appl. 6, e16253 (2017). [CrossRef]   [PubMed]  

7. N. Singh, D. D. Hudson, Y. Yu, C. Grillet, S. D. Jackson, A. Casas-Bedoya, A. Read, P. Atanackovic, S. G. Duvall, S. Palomba, B. Luther-Davies, S. Madden, D. J. Moss, and B. J. Eggleton, “Midinfrared supercontinuum generation from 2 to 6 μm in a silicon nanowire,” Optica 2, 797–802 (2015). [CrossRef]  

8. A. V. Krasavin, P. Ginzburg, G. A. Wurtz, and A. V. Zayats, “Nonlocality-driven supercontinuum white light generation in plasmonic nanostructures,” Nat. Commun. 7, 11497 (2016). [CrossRef]   [PubMed]  

9. X. Guo, C. L. Zou, C. Schuck, H. Jung, R. Cheng, and H. X. Tang, “Parametric down-conversion photon-pair source on a nanophotonic chip,” Light. Sci. Appl. 6, e16249 (2017). [CrossRef]   [PubMed]  

10. G. Harder, T. J. Bartley, A. E. Lita, S. W. Nam, T. Gerrits, and C. Silberhorn, “Single-mode parametric-down-conversion states with 50 photons as a source for mesoscopic quantum optics,” Phys. Rev. Lett. 116, 143601 (2016). [CrossRef]   [PubMed]  

11. X. Guo, C. L. Zou, and H. X. Tang, “Second-harmonic generation in aluminum nitride microrings with 2500%/w conversion efficiency,” Optica 3, 1126–1131 (2016). [CrossRef]  

12. M. Allgaier, V. Ansari, L. Sansoni, C. Eigner, V. Quiring, R. Ricken, G. Harder, B. Brecht, and C. Silberhorn, “Highly efficient frequency conversion with bandwidth compression of quantum light,” Nat. Commun. 8, 14288 (2017). [CrossRef]   [PubMed]  

13. C. Joshi, A. Farsi, S. Clemmen, S. Ramelow, and A. L. Gaeta, “Frequency multiplexing for quasi-deterministic heralded single-photon sources,” Nat. Commun. 9, 847 (2018). [CrossRef]   [PubMed]  

14. D. T. Spencer, T. Drake, T. C. Briles, J. Stone, L. C. Sinclair, C. Fredrick, Q. Li, D. Westly, B. R. Ilic, A. Bluestone, N. Volet, T. Komljenovic, L. Chang, S. H. Lee, D. Y. Oh, M.-G. Suh, K. Y. Yang, M. H. P. Pfeiffer, T. J. Kippenberg, E. Norberg, L. Theogarajan, K. Vahala, N. R. Newbury, K. Srinivasan, J. E. Bowers, S. A. Diddams, and S. B. Papp, “An optical-frequency synthesizer using integrated photonics,” Nature 557, 81–85 (2018). [CrossRef]   [PubMed]  

15. A. Rueda, F. Sedlmeir, M. C. Collodo, U. Vogl, B. Stiller, G. Schunk, D. V. Strekalov, C. Marquardt, J. M. Fink, O. Painter, G. Leuchs, and H. G. L. Schwefel, “Efficient microwave to optical photon conversion: an electro-optical realization,” Optica 3, 597–604 (2016). [CrossRef]  

16. S. Sederberg and A. Y. Elezzabi, “Nonmonotonic wavelength-dependent power scaling in silicon-on-insulator waveguides via nonlinear optical effect conglomeration,” ACS Photonics 1, 576–581 (2014). [CrossRef]  

17. A. Levanon, A. Dahan, A. Nagler, E. Lifshitz, E. Bahar, M. Mrejen, and H. Suchowski, “Pulse shaping of broadband adiabatic shg from a Ti-sapphire oscillator,” Opt. Lett. 42, 2992–2995 (2017). [CrossRef]   [PubMed]  

18. H. Suchowski, G. Porat, and A. Arie, “Adiabatic processes in frequency conversion,” Laser Photon. Rev. 8, 333–367 (2014). [CrossRef]  

19. T. Suhara and H. Nishihara, “Theoretical analysis of waveguide second-harmonic generation phase matched with uniform and chirped gratings,” IEEE J. Quantum Electron. 26, 1265–1276 (1990). [CrossRef]  

20. J. B. Driscoll, N. Ophir, R. R. Grote, J. I. Dadap, N. C. Panoiu, K. Bergman, and R. M. Osgood, “Width-modulation of si photonic wires for quasi-phase-matching of four-wave-mixing: experimental and theoretical demonstration,” Opt. Express 20, 9227–9242 (2012). [CrossRef]   [PubMed]  

21. B. Y. Jin, J. H. Yuan, C. X. Yu, X. Z. Sang, S. Wei, X. T. Zhang, Q. Wu, and G. Farrell, “Efficient and broadband parametric wavelength conversion in a vertically etched silicon grating without dispersion engineering,” Opt. Express 22, 6257–6268 (2014). [CrossRef]   [PubMed]  

22. S. Richard, “Second-harmonic generation in tapered optical fibers,” J. Opt. Soc. Am. B 27, 1504–1512 (2010). [CrossRef]  

23. R. Prinja, “Broadband optical three wave mixing in algaas waveguides,” Ph.D. thesis (2015).

24. W. H. Pernice, C. Xiong, C. Schuck, and H. X. Tang, “Second harmonic generation in phase matched aluminum nitride waveguides and micro-ring resonators,” Appl. Phys. Lett. 100, 223501 (2012). [CrossRef]  

25. X. Guo, C. L. Zou, H. Jung, and H. X. Tang, “On-chip strong coupling and efficient frequency conversion between telecom and visible optical modes,” Phys. Rev. Lett. 117, 123902 (2016). [CrossRef]   [PubMed]  

26. R. W. Boyd, Nonlinear Optics (Academic Press, 2003).

27. K. Bergmann, H. Theuer, and B. W. Shore, “Coherent population transfer among quantum states of atoms and molecules,” Rev. Mod. Phys. 70, 1003 (1998). [CrossRef]  

28. M. Mojahedi, K. Malloy, G. Eleftheriades, J. Woodley, and R. Chiao, “Abnormal wave propagation in passive media,” IEEE J. Sel. Top. Quantum Electron. 9, 30–39 (2003). [CrossRef]  

29. L. Chang, A. Boes, X. Guo, D. T. Spencer, M. J. Kennedy, J. D. Peters, N. Volet, J. Chiles, A. Kowligy, N. Nader, D. D. Hickstein, E. J. Stanton, S. A. Diddams, S. B. Papp, and J. E. Bowers, “Heterogeneously integrated gaas waveguides on insulator for efficient frequency conversion,” Laser Photon. Rev. 12, 1800149 (2018). [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 (7)

Fig. 1
Fig. 1 (a) Schematic illustration of the tapered waveguide deposited on silica buffer whose width varies along the light propagation direction as w(z), with h = 300 nm. Insets: typical eigenmode distributions for signal, pump, and idler photons. (b) The effective refractive indices neff of eigenmodes at signal (green), pump (red), and idler (blue) wavelengths as a function of waveguide width w. Gray dashed line indicates the point where the phase matching condition is satisfied.
Fig. 2
Fig. 2 (a) Nonlinear conversion spectra of λ3 for δw = 0, 4, 8 nm, L = 10000 μm, and P2 = 1 W. (b) Bandwidth Δλ extracted from conversion spectra as a function of δw.
Fig. 3
Fig. 3 (a) Integral of η over the entire spectrum as a function of δw for L = 1000 μm, L = 10000 μm, and P2 = 1 W. (b) Integral conversion efficiency as a function of P2 for δw = 0 and δw = 4 nm with L = 1000 μm. Light blue rectangle highlights the parameter range where the “area law” holds (P2 < 6 W).
Fig. 4
Fig. 4 (a) Conversion efficiency η of idler photons as a function of signal λ1 (x-axis) and pump λ2 (y-axis) wavelengths, with L = 10000 μm and δw = 0 nm. (b) Conversion efficiency η of idler photons as a function of signal λ1 (x-axis) and pump λ2 (y-axis) wavelengths, with L = 10000 μm and δw = 4 nm.
Fig. 5
Fig. 5 The phase spectrum of the complex conversion coefficient for different waveguide width difference δw, with L = 1000 μm and P2 = 1 W.
Fig. 6
Fig. 6 (a) The frequency conversion spectrum of idler photons for different waveguide width deviation Δw, with δw = 4nm, L = 10000μm, and P2 = 1W. (b) The frequency conversion efficiency at 600.4 nm as a function of Δw for different waveguide width difference δw, with L = 10000μm and P2 = 1W.
Fig. 7
Fig. 7 The frequency conversion efficiency under pump power of P2 = 1W and P2 = 5W, for uniform waveguide (a) and tapered waveguide (b), respectively.

Equations (16)

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

= ( Δ β 2 g * g Δ β 2 ) .
i d d z | Φ ( z ) = ( z ) | Φ ( z ) ,
| Φ ( z ) = | Φ ( 0 ) = 𝒯 e i 0 z ( s ) d s | Φ ( 0 ) .
Δ λ δ w d ( Δ β ) d w / d ( Δ β ) d λ ,
η d λ 2 π | g | 2 L / d ( Δ β ) d λ .
η = 1 e 2 π | g | 2 | d Δ β / d z | 2 π | g | 2 | d Δ β / d z | ,
η d λ 2 π | g | 2 d ( Δ β ) d λ d z = 2 π | g | 2 L / d ( Δ β ) d λ ,
η = | g | 2 δ z 2 sinc 2 ( | g | δ z ) ,
d λ d w = Δ λ δ w .
Δ λ = δ w d λ d w = δ ω 1 / d w 1 / d λ = δ ω d ( Δ β ) d w / d ( Δ β ) d λ .
Δ β ( λ 1 , λ 2 , λ 3 ) = 2 π ( n 1 λ 1 + n 2 λ 2 n 3 λ 3 ) .
2 π c λ 1 + 2 π c λ 2 = 2 π c λ 3 ,
Δ β ( λ 2 , λ 3 ) = 2 π ( n 2 n 1 λ 2 n 3 n 1 λ 3 ) .
d ( Δ β ) d λ 3 = 2 π ( n 3 n 1 ) λ 3 2 .
η d λ 2 π | g | 2 | d ( Δ β ) / d z | d λ = 2 π | g | 2 | d ( Δ β ) / d λ | d z = 2 π | g | 2 L / d ( Δ β ) d λ ,
w ( z ) = ( w 0 + Δ w ) + δ w L ( z L 2 ) ,
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.