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

Simple method for locking birefringent resonators

Open Access Open Access

Abstract

We report on a simple method of locking a laser to a birefringent cavity using polarization spectroscopy. The birefringence of the resonator permits the simple extraction of an error signal by using one polarization state as a phase reference for another state. No modulation of the light or the resonator is required, reducing the complexity of the laser locking setup. This method of producing an error signal can be used on most birefringent optical resonators, even if the details of birefringence and eigenpolarizations are not known. This technique is particularly well suited for fiber ring resonators due to the inherent birefringence of the fiber and the unknown nature of that birefringence. We present an experimental demonstration of this technique using a fiber ring.

© 2015 Optical Society of America

1. Introduction

Using optical resonators to measure and control the frequency of a laser is a well-established technique. The use of locking methods in which multiple modes of a probe light field are compared are especially effective. Examples include the Pound-Drever-Hall (PDH) locking scheme [1, 2], in which different frequency components are interfered, and schemes in which different spatial modes are compared [3, 4]. In general, any two orthogonal modes can be used to make a phase-sensitive error signal [5]. In this paper, we generalize the locking technique of Hänsch and Couillaud (HC) [6], which utilizes different polarization modes to produce an error signal. The HC method uses a polarizer in the resonator to produce a polarization-dependent resonance condition. Variations of this technique have been demonstrated in free space cavities by introducing a birefringent crystal into the cavity [7], using a non-planar ring cavity [8], using a triangular cavity [9], or using the birefringence of dielectric mirrors [10]. A special case of this technique has been demonstrated in fiber for strain sensing applications [11]. We generalize these locking techniques and describe and demonstrate a polarization spectroscopy locking method that introduces no losses, is simple to implement, and requires no knowledge of the cavity birefringence. These properties make this method particularly attractive for systems with inherent or unknown birefringence, such as whispering-gallery-mode (WGM) resonators [12, 13], cavities formed with crystal-coated mirrors [14, 15], and fiber-based applications.

The resonance properties [16] and the polarization effects of fiber rings have been investigated previously [1719]. Small stresses and imperfections in the core of the fiber produce polarization-dependent phase shifts that are rarely known a priori. This inherent unknown birefringence make this locking technique well suited to use with fiber ring resonators. In addition to the ease of alignment and mode-matching that comes with using single-mode optical fibers, fiber rings allow for long cavities and narrow linewidths in compact packages with minimal back reflection to the laser source. Traditional methods such as PDH can be used to lock to fiber rings [20], but there are techniques unique to fiber optics, such as using Rayleigh scattering to provide optical feedback [21]. In contrast, the locking scheme we present requires no modulation or demodulation, which allows for simple electronics and a high-bandwidth error signal. Further, it avoids the challenges of optical feedback. We demonstrate this technique experimentally with a fiber ring. This method can be used to produce compact pre-stabilized lasers for use in fiber-based systems that require stable coherent sources, such as telecommunication systems [22], lidar [23], fiber gyroscopes [24], and other fiber-based sensors, such as strain sensors [11].

2. Polarization in optical resonators

Most optical resonators will have some effect on the polarization of circulating light. A birefringent cavity will impart a different phase delay to each polarization and may alter the polarization of the input light. Additionally, cavity losses may be polarization dependent, which can create additional interesting polarization effects.

These polarization effects can be more easily understood by representing the electric field of fully polarized light as a two-component complex vector known as a Jones vector [25]. Conventionally, this vector is written using linear horizontal and vertical polarizations as the basis, so a polarized plane light wave at time t a distance z along the propagation axis will have the form

E=(EHEV)=(AHAVeiϕ)ei(ωtkz),
where AH and AV are the amplitudes of each polarization component, ϕ is the phase difference between each component, and ω and k are the angular frequency and wavenumber of the light, respectively. The last exponential is usually dropped since global phases have no effect on the polarization. In this framework, cavity losses, polarization transformations, and phase changes can be described by a 2 × 2 Jones matrix that acts on the polarization vector. For example, in this basis the Jones matrix for a half-wave plate with fast axis at angle θ to the horizontal is
(cos2θsin2θsin2θcos2θ).

In what follows, we will consider the polarization effects of a fiber ring resonator with a single input coupler, but this method is completely generalizable to any birefringent cavity.

The amplitude transmittance and reflectance matrices of the input coupler are labeled 𝕋 and ℝ, respectively. Let 𝔽 be the matrix that represents the polarization transformation as the light propagates in the cavity (see Fig. 1). Included in this matrix is information about the overall (frequency-dependent) phase picked up and any losses incurred as light propagates from the input coupler through the cavity and back. The total cavity amplitude reflectivity matrix ℝcav, defined by E⃗refl = ℝcavE⃗in, is given by

cav=𝕋𝔽(𝕀𝔽)1𝕋,
where 𝕀 is the identity matrix. Since ℝ𝔽 represents one round trip of the resonator, the two eigenvectors of this matrix are the eigenpolarizations of the resonator. For positive real eigenvalues of ℝ𝔽, light will resonate.

 figure: Fig. 1

Fig. 1 Block diagram of an optical experiment compatible with our locking method. Polarization controllers (PCs) on the input and output of the resonator are given by the Jones matrices 𝕐in and 𝕐out. Matrices ℝ and 𝕋 represent reflection and transmission of the coupler, respectively, and 𝔽 represents propagation through the resonator. A polarizing beam splitter (PBS) follows at the end.

Download Full Size | PDF

We assume that the eigenvectors of ℝ𝔽 are also eigenvectors of 𝕋 and ℝ, and therefore of ℝcav, for all frequencies ω. This assumption holds for input couplers with polarization-independent properties. For birefringent input couplers, this assumption is valid when the eigen-polarizations of the coupler match those of the cavity. For fibers, the polarization effects of the couplers are often much smaller than the effects due to birefringence in the fiber itself [26], so this is a reasonable assumption for fiber rings.

In this case, the eigenvalues of ℝcav are given by

rcavj(ω)=rjtj2fj(ω)1rjfj(ω),
where tj, rj, and fj are the eigenvalues of eigenpolarization j under 𝕋, ℝ, and 𝔽, respectively. In general, fj = αjeikjp, where kj = 2πnjν/c is the wavenumber for each eigenpolarization, nj is the effective index of refraction for each eigenpolarization, ν is the frequency of light, p is the distance the light travels in one round trip of the cavity, and αj is a complex term that accounts for losses in the cavity and additional phase shifts not due to propagation. Near resonance, the reflected light undergoes a very large phase shift. Figure 2 displays the real and imaginary parts of this reflection coefficient rcavj(ω) for a general optimally coupled cavity. Note the sharp change in the imaginary component near resonance due to the phase shift.

 figure: Fig. 2

Fig. 2 Cavity reflectance rcav as a function of frequency for a single input mode. Blue/Solid: Real, Red/Dashed: Imaginary. Input mirror reflectivity r2 = 1 − t2 = 0.95, cavity losses (1 − α2) = 0.05. Inset: closeup of resonance peak. Note the sharp change in the imaginary component as the frequency moves through a cavity resonance.

Download Full Size | PDF

In general, the wavenumbers for different eigenpolarizations will not be equal, so there will be a difference in the phase accumulated by each eigenpolarization after one round trip, resulting in two longitudinal modes that are on resonance for different ω. As one eigenpolarization moves through a resonance, it undergoes a large phase shift on reflection, while the other eigenpolarization does not. In this manner, the latter can be used as a phase reference for the former. This is easiest to implement when the resonances are shifted sufficiently far in frequency. Individual resonance peaks are well separated for round trip phase differences between eigenpolarizations Δθ that satisfy

Δθmod2π>2πΔνFSR
and
Δθmod2π>2πΔνFSR,
where Δν is the full width at half maximum of the cavity resonance and FSR is the free spectral range. This condition can be satisfied even in cavities with small birefringence so long as the cavity has a large finesse ( FSRΔν).

3. The error signal

To use the birefringence of the resonator to produce an error signal, light traveling to the resonator must first be put into the correct input polarization, which can be done using a polarization controller (PC). There are many ways to implement such a device, but a common arrangement consists of a quarter-wave plate, a half-wave plate, and a second quarter-wave plate, each of which can be rotated independently. These polarization controllers can map an arbitrary input polarization state to any other polarization state [27]. After reflecting from the resonator, light goes through another PC before propagating to a polarizing beam splitter (PBS) with a photodiode (PD) at each output (Fig. 1). With the correct polarization control settings before and after the cavity, the difference in power at the two output ports of the PBS produces an error signal for the resonance condition of the cavity. Expressed in the Jones matrix formalism the error signal Δ is proportional to

Δ|Eout,2|2|Eout,1|2,
where
Eout,1=(10)𝕐outcav𝕐inElaserEout,2=(01)𝕐outcav𝕐inElaser
are the respective projections of the final electric field amplitude onto the PBS polarization basis states. The proportionality constant will depend on the PD gain. In the above equations, 𝕐in (out) is the Jones matrix of the input (output) PC and ℝcav is the cavity amplitude reflectivity matrix given by Eq. (3).

We further assume the cavity eigenpolarizations are orthogonal. This will be true for cavities with polarization-independent losses, in which ℝ, 𝕋, and 𝔽 are all unitary matrices multiplied by a constant loss term [18]. The eigenpolarizations will also be orthogonal for cavities in which one polarization mode is completely extinguished in the cavity, as in the original Hänsch-Couillaud method [6]. In practice, this approximation holds true for most single-mode fiber rings [18].

Let E⃗a and E⃗b be the normalized eigenpolarizations of ℝ𝔽. The input PCs are set so that the input polarization is an equal superposition of these eigenpolarizations:

Ein=𝕐inElaser=E02(Ea+eiγEb),
where E0 is the amplitude of the electric field and γ is the phase difference between eigenpolarization components. The reflected light will be in the polarization state
Erefl=cavEin=E02(cavEa+eiγcavEb)=E02(rcava(ω)Ea+eiγrcavb(ω)Eb).
Because the polarization controllers can map an arbitrary input state to any given output polarization, there exists an arrangement of the output polarization controller that maps one eigenpolarization to an equal superposition of the PBS polarizations:
Ea𝕐outEa=12(1eiδ)
in the PBS basis for some δ. The polarization controller produces a lossless, and therefore unitary, transformation on the Jones vector, so the other eigenpolarization will be mapped to a final polarization orthogonal to this with some relative phase shift ϕ:
Eb𝕐outEb=eiϕ2(1eiδ).
The final polarization state in the PBS basis is then
Eout=𝕐outErefl=E02(rcava(ω)(1eiδ)+ei(γ+ϕ)rcavb(ω)(1eiδ))
and the sum and difference of the photodiode signals are proportional to
|Eout,1|2+|Eout,2|2=E022(|rcava(ω)|2+|rcavb(ω)|2)
and
|Eout,2|2|Eout,1|2=E02Re{(rcava(ω))*rcavb(ω)ei(γ+ϕ)},
respectively. Re denotes the real part and represents complex conjugation. Equation (15) forms the error signal.

In the case of widely separated resonances, near the resonance of one eigenpolarization the imaginary part of the cavity reflectivity undergoes a sharp change, while the reflectivity for the other eigenpolarization changes very little. By adjusting the additional phase term in Eq. (15), the steep imaginary part of the individual reflection coefficients can be extracted. For widely separated resonances, this occurs for γ + ϕ ≈ ±π/2. The output polarization controllers vary ϕ to produce an ideal error signal; therefore, it is not necessary to know or control γ. This means that the input polarization state can be chosen without regard to the relative phases of the eigenpolarizations, and that only the sum signal is needed to set the input PCs.

This is a general method for producing an error signal for a birefringent resonator with two longitudinal modes; for a chosen input polarization satisfying Eq. (9), one can always find a physically realizable Jones matrix 𝕐out that produces an error signal.

Fundamentally, this method is limited only by shot noise, which scales as P−1/2, where P is the light incident on the cavity. In fiber systems, stimulated Brillouin scattering [28] puts a limit on the input power, which prevents the reduction of shot noise by turning up the optical power. However, in practice this method will be limited by acoustic/vibrational noise and thermorefractive effects in the fiber.

4. Experimental design and data

The experimental setup needed to demonstrate this method is shown in Fig. 3. We use a fiber ring resonator as the birefringent cavity. This trivializes cavity alignment and mode-matching. Birefringence in optical fibers is often viewed as a trait that must be overcome [17], but here we use it as a feature and exploit it to produce the error signal. We use a direct-coupling fiber ring setup [16] described below to produce the error signal.

 figure: Fig. 3

Fig. 3 A schematic of the experimental setup used to produce the error signal. The fiber polarization controllers (FPCs) approximate adjustable waveplates to tune the input and output polarizations to optimize the error signal. The 20-m fiber ring has a finesse of ≈ 60. The sum signal is used to ensure equal input power in each eigenpolarization.

Download Full Size | PDF

A non-planar ring oscillator (NPRO) Nd:YAG laser emitting at 1064 nm (Lightwave 125-1064-700) provides the incident light and has an observed free-running frequency noise of 100Hz/Hz at 100Hz and 1/f frequency dependence. The output frequency can be controlled by changing the temperature of the lasing medium crystal or by applying a strain to the crystal via a piezoelectric actuator. Approximately 1 mW of linearly-polarized light, with the power selected to avoid stimulated Brillouin scattering in the fiber ring resonator, is coupled into a single-mode (SM) optical fiber and encounters the first fiber polarization controller (FPC). Our controller consists of three rotating paddles with varying lengths of fiber coiled around each paddle. This approximates a quarter-wave plate, followed by a half-wave plate, and finally another quarter-wave plate. Using this setup, we are able to map any input polarization to an arbitrary output polarization.

From there, the light enters an SM 95:5 2 × 2 coupler spliced to a p = 20 m length of single-mode fiber, forming a ring. This coupling ratio was chosen to produce a nearly optimally coupled resonator, since we expect approximately 4% total losses in the splices and in the fiber itself. The losses and coupling ratio result in a finesse of approximately 60 for both eigenpolarizations and a free spectral range of c/np ≈ 10 MHz [29] for a fiber core with index n ≈ 1.46. Note that an error signal can still be produced in a resonator that is not optimally coupled, though the error signal may not be as steep. The ring is placed in a box that provides acoustic and thermal isolation.

Following the coupler, the light again goes through an FPC, and then on to a fiber PBS. Each output of the PBS goes to a photodiode, the outputs of which are subtracted using a low noise amplifier to produce the error signal.

To optimize the error signal, the input polarization must be an equal superposition of the cavity eigenpolarizations. This is achieved by scanning the laser frequency and adjusting the input FPC until the resonant peaks in the total reflection signal—obtained from the sum of the PD signals—are balanced, as shown in Fig. 4(a).

 figure: Fig. 4

Fig. 4 Observed and modeled (a) Sum and (b) difference of the two output ports of the PBS for the fiber ring resonator setup. The modeled resonator has a 95:5 coupler with 2% loss and the fiber has 4% loss. (a) Input power is split between cavity eigenpolarizations. (b) The output polarization controller was set to optimize the error signal. Discrepancies between theory and data are likely due to low-passing in the photodiode.

Download Full Size | PDF

The last step is to adjust the output polarization control paddles to produce an error signal matching Fig. 4(b). This is done empirically; no calculation was required to produce the error signal. Looking at the difference signal of the photodiodes, the polarization controllers are adjusted to make the peaks in the error signal the same amplitude for both resonances, and to make the signal symmetric about the frequencies which are halfway between resonances.

A typical reflected power and error signal for the 20-m fiber ring setup is shown in Fig. 4. Note the agreement with the theoretical model presented in section 3. Once the losses and coupler properties are measured, the only free parameter fit in the theoretical model is Δθ. It is important to note that we do not need to know what the eigenpolarizations of the resonator are, or which particular input polarization (satisfying Eq. (9)) is being used to produce this error signal.

This error signal is stable for longer than 24 hours with the same input and output polarization controller positions, demonstrating the robustness of the system to slow fluctuations in the laboratory environment. We have measured the noise spectrum of the resonator error signal and find it to be suitable for laser frequency stabilization even for an NPRO, one of the lowest-noise lasers commercially available (paper in preparation).

5. Conclusions

We have presented a simple yet general method for producing an error signal for a birefringent resonator. This method relies on the phase difference acquired by the different eigenpolarizations of the resonator on reflection, where light in one eigenpolarization is used as a phase reference for light in the other. The method does not require any particular arrangement of eigenpolarizations, or even that they be known to the experimenter. Instead, all that is required is the ability to tune both the polarization of the light incident on the resonator and the polarization of the reflected light before splitting it on a PBS. Since the generation of this error signal does not require any modulation of the light or of the cavity, the electronics used to produce the error signal are remarkably simple. Further, because there is no demodulation, a low-pass filter is not needed to extract the error signal, resulting in a high-bandwidth error signal, a requirement for fast feedback. We have demonstrated this method experimentally by producing an error signal with a 20-m long fiber ring resonator. While this method can be applied to nearly any birefringent optical cavity, this technique is especially useful for fiber-based sensors. This method can be used to produce compact, pre-stabilized lasers, with particular relevance for fiber-based laser systems.

References and links

1. R. V. Pound, “Electronic frequency stabilization of microwave oscillators,” Rev. Scientific Instrum. 17, 490–505 (1946). [CrossRef]  

2. R. W. P. Drever, J. L. Hall, F. V. Kowalski, J. Hough, G. M. Ford, A. J. Munley, and H. Ward, “Laser phase and frequency stabilization using an optical resonator,” Appl. Phys. B 31, 97–105 (1983). [CrossRef]  

3. D. A. Shaddock, M. B. Gray, and D. E. McClelland, “Frequency locking a laser to an optical cavity by use of spatial mode interference,” Opt. Lett. 24, 1499–1501 (1999). [CrossRef]  

4. J. Miller and M. Evans, “Length control of an optical resonator using second-order transverse modes,” Opt. Lett. 39, 2495–2498 (2014). [CrossRef]   [PubMed]  

5. M. D. Harvey and A. G. White, “Frequency locking by analysis of orthogonal modes,” Opt. Commun. 221, 163–171 (2003). [CrossRef]  

6. T. Hänsch and B. Couillaud, “Laser frequency stabilization by polarization spectroscopy of a reflecting reference cavity,” Opt. Commun. 35, 441–444 (1980). [CrossRef]  

7. J. Boon-Engering, W. van der Veer, E. Bente, and W. Hogervorst, “Stabilization of an optical cavity containing a birefringent element,” Opt. Commun. 140, 285–288 (1997). [CrossRef]  

8. Y. Honda, H. Shimizu, M. Fukuda, T. Omori, J. Urakawa, K. Sakaue, H. Sakai, and N. Sasao, “Stabilization of a non-planar optical cavity using its polarization property,” Opt. Commun. 282, 3108–3112 (2009). [CrossRef]  

9. S. Moriwaki, T. Mori, K. Takeno, and N. Mio, “Frequency discrimination method making use of polarization selectivity of triangular optical cavity,” Appl. Phys. Express 2, 016501 (2009). [CrossRef]  

10. P. Asenbaum and M. Arndt, “Cavity stabilization using the weak intrinsic birefringence of dielectric mirrors,” Opt. Lett. 36, 3720–3722 (2011). [CrossRef]   [PubMed]  

11. G. Gagliardi, S. De Nicola, P. Ferraro, and P. De Natale, “Interrogation of fiber Bragg-grating resonators by polarization-spectroscopy laser-frequency locking,” Opt. Express 15, 3715–3728 (2007). [CrossRef]   [PubMed]  

12. D. V. Strekalov, R. J. Thompson, L. M. Baumgartel, I. S. Grudinin, and N. Yu, “Temperature measurement and stabilization in a birefringent whispering gallery mode resonator,” Opt. Express 19, 14495–14501 (2011). [CrossRef]   [PubMed]  

13. A. Schliesser, R. Rivière, G. Anetsberger, O. Arcizet, and T. J. Kippenberg, “Resolved-sideband cooling of a micromechanical oscillator,” Nature Phys. 4, 415–419 (2008). [CrossRef]  

14. G. D. Cole, W. Zhang, M. J. Martin, J. Ye, and M. Aspelmeyer, “Tenfold reduction of Brownian noise in high-reflectivity optical coatings,” Nature Photon. 7, 644–650 (2013). [CrossRef]  

15. T. Kessler, C. Hagemann, C. Grebing, T. Legero, U. Sterr, F. Riehle, M. J. Martin, L. Chen, and J. Ye, “A sub-40-mHz-linewidth laser based on a silicon single-crystal optical cavity,” Nature Photon. 6, 687–692 (2012). [CrossRef]  

16. F. Zhang and J. W. Y. Lit, “Direct-coupling single-mode fiber ring resonator,” J. Opt. Soc. Am. A 5, 1347–1355 (1988). [CrossRef]  

17. B. Lamouroux, B. Prade, and A. Orszag, “Polarization effect in optical-fiber ring resonators,” Opt. Lett. 7, 391–393 (1982). [CrossRef]   [PubMed]  

18. K. Iwatsuki, K. Hotate, and M. Higashiguchi, “Eigenstate of polarization in a fiber ring resonator and its effect in an optical passive ring-resonator gyro,” Appl. Opt. 25, 2606–2612 (1986). [CrossRef]   [PubMed]  

19. W. Wang and J. Xia, “The characterizations of polarization in resonator integrated optic gyroscope,” Opt. Quantum Electron. 42, 313–325 (2011). [CrossRef]  

20. P.-H. Merrer, O. Llopis, and G. Cibiel, “Laser stabilization on a fiber ring resonator and application to RF filtering,” IEEE Photon. Technol. Lett. 20, 1399–1401 (2008). [CrossRef]  

21. T. J. Paul and E. A. Swanson, “Narrow-linewidth diode laser using a fiber-optic ring resonator,” Opt. Lett. 18, 1241–1243 (1993). [CrossRef]   [PubMed]  

22. W. F. Brinkman, T. L. Koch, D. V. Lang, and D. P. Wilt, “The lasers behind the communications revolution,” Bell Labs Techn. J. 5, 150–167 (2002). [CrossRef]  

23. M. Ostermeyer, P. Kappe, R. Menzel, and V. Wulfmeyer, “Diode-pumped Nd:YAG master oscillator power amplifier with high pulse energy, excellent beam quality, and frequency-stabilized master oscillator as a basis for a next-generation lidar system,” Appl. Opt. 44, 582–590 (2005). [CrossRef]   [PubMed]  

24. H. C. Lefevre, R. a. Bergh, and H. J. Shaw, “All-fiber gyroscope with inertial-navigation short-term sensitivity,” Opt. Lett. 7, 454–456 (1982). [CrossRef]   [PubMed]  

25. E. Hecht, Optics (Pearson Education, 2008).

26. W. Burns, “Polarization characteristics of single-mode fiber couplers,” IEEE Trans. Microwave Theory Tech. 30, 1577–1588 (1982). [CrossRef]  

27. Z. Xiao-Guang and Z. Yuan, “The number of least degrees of freedom required for a polarization controller to transform any state of polarization to any other output covering the entire Poincaré sphere,” Chin. Phys. B 17, 2509–2513 (2008). [CrossRef]  

28. E. Ippen, “Stimulated Brillouin scattering in optical fibers,” Appl. Phys. Lett. 21, 539–541 (1972). [CrossRef]  

29. A. E. Siegman, Lasers (University Science Books, 1986).

Cited By

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

Alert me when this article is cited.


Figures (4)

Fig. 1
Fig. 1 Block diagram of an optical experiment compatible with our locking method. Polarization controllers (PCs) on the input and output of the resonator are given by the Jones matrices 𝕐in and 𝕐out. Matrices ℝ and 𝕋 represent reflection and transmission of the coupler, respectively, and 𝔽 represents propagation through the resonator. A polarizing beam splitter (PBS) follows at the end.
Fig. 2
Fig. 2 Cavity reflectance rcav as a function of frequency for a single input mode. Blue/Solid: Real, Red/Dashed: Imaginary. Input mirror reflectivity r2 = 1 − t2 = 0.95, cavity losses (1 − α2) = 0.05. Inset: closeup of resonance peak. Note the sharp change in the imaginary component as the frequency moves through a cavity resonance.
Fig. 3
Fig. 3 A schematic of the experimental setup used to produce the error signal. The fiber polarization controllers (FPCs) approximate adjustable waveplates to tune the input and output polarizations to optimize the error signal. The 20-m fiber ring has a finesse of ≈ 60. The sum signal is used to ensure equal input power in each eigenpolarization.
Fig. 4
Fig. 4 Observed and modeled (a) Sum and (b) difference of the two output ports of the PBS for the fiber ring resonator setup. The modeled resonator has a 95:5 coupler with 2% loss and the fiber has 4% loss. (a) Input power is split between cavity eigenpolarizations. (b) The output polarization controller was set to optimize the error signal. Discrepancies between theory and data are likely due to low-passing in the photodiode.

Equations (15)

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

E = ( E H E V ) = ( A H A V e i ϕ ) e i ( ω t k z ) ,
( cos 2 θ sin 2 θ sin 2 θ cos 2 θ ) .
cav = 𝕋 𝔽 ( 𝕀 𝔽 ) 1 𝕋 ,
r cav j ( ω ) = r j t j 2 f j ( ω ) 1 r j f j ( ω ) ,
Δ θ mod 2 π > 2 π Δ ν FSR
Δ θ mod 2 π > 2 π Δ ν FSR ,
Δ | E out , 2 | 2 | E out , 1 | 2 ,
E out , 1 = ( 1 0 ) 𝕐 out cav 𝕐 in E laser E out , 2 = ( 0 1 ) 𝕐 out cav 𝕐 in E laser
E in = 𝕐 in E laser = E 0 2 ( E a + e i γ E b ) ,
E refl = cav E in = E 0 2 ( cav E a + e i γ cav E b ) = E 0 2 ( r cav a ( ω ) E a + e i γ r cav b ( ω ) E b ) .
E a 𝕐 out E a = 1 2 ( 1 e i δ )
E b 𝕐 out E b = e i ϕ 2 ( 1 e i δ ) .
E out = 𝕐 out E refl = E 0 2 ( r cav a ( ω ) ( 1 e i δ ) + e i ( γ + ϕ ) r cav b ( ω ) ( 1 e i δ ) )
| E out , 1 | 2 + | E out , 2 | 2 = E 0 2 2 ( | r cav a ( ω ) | 2 + | r cav b ( ω ) | 2 )
| E out , 2 | 2 | E out , 1 | 2 = E 0 2 Re { ( r cav a ( ω ) ) * r cav b ( ω ) e i ( γ + ϕ ) } ,
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.