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

Thermal analysis of optical reference cavities for low sensitivity to environmental temperature fluctuations

Open Access Open Access

Abstract

The temperature stability of optical reference cavities is significant in state-of-the-art ultra-stable narrow-linewidth laser systems. In this paper, the thermal time constant and thermal sensitivity of reference cavities are analyzed when reference cavities respond to environmental perturbations via heat transfer of thermal conduction and thermal radiation separately. The analysis as well as simulation results indicate that a reference cavity enclosed in multiple layers of thermal shields with larger mass, higher thermal capacity and lower emissivity is found to have a larger thermal time constant and thus a smaller sensitivity to environmental temperature perturbations. The design of thermal shields for reference cavities may vary according to experimentally achievable temperature stability and the coefficient of thermal expansion of reference cavities. A temperature fluctuation-induced length instability of reference cavities as low as 6 × 10−16 on a day timescale can be achieved if a two-layer thermal shield is inserted between a cavity with the coefficient of thermal expansion of 1 × 10−10 /K and an outer vacuum chamber with temperature fluctuation amplitude of 1 mK and period of 24 hours.

© 2015 Optical Society of America

1. Introduction

Optical reference cavities with finesse of >100,000 are widely used to stabilize the frequency of lasers in optical atomic clocks, precision metrology, high resolution spectroscopy, tests of fundamental physics and deep space navigation [17], which rely on the high frequency stability of lasers. When a laser is tightly frequency-stabilized to a reference cavity using the Pound-Drever-Hall technique [8], the laser frequency stability is dominated by the effective length stability of the reference cavity. To minimize the length fluctuation, reference cavities are often made of ultra-low expansion (ULE) glass or single crystal silicon, and are installed in vacuum chambers with high stability temperature stabilization for less sensitivity to environmental temperature fluctuations via thermal expansion [912]. Special geometry and mounting configurations for reference cavities have been developed for less sensitivity to environmental vibration [1317]. When stabilized to those carefully designed reference cavities, lasers with a frequency instability at the 10−16 level at 1–10 s averaging time have been achieved [11, 12, 14, 18, 19], approaching the thermal noise limit of reference cavities [20].

However, for the applications of ultra-stable laser systems, both the short-term stability on a few seconds’ timescale as well as the long-term stability over a day are essential. In most of those ultra-stable laser systems, the laser frequency drifts linearly at tens of mHz/s to several Hz/s on a relatively short timescale, and drifts nonlinearly on hundreds of seconds’ timescale which is hardly compensated by applying a feed forward frequency correction. The aging of the cavity material and optical contact between cavity mirrors and spacers leads to a fractional length change on the order of 10−17/s to 10−16/s [21]. Therefore, temperature fluctuation of reference cavities is the key cause of laser frequency drift. Well-designed temperature controllers can achieve sub-mK stability over several days [10,22]. When a reference cavity made of ULE glass is temperature-stabilized at the zero-crossing thermal expansion temperature (TCTE=0) with accuracy of 0.1 °C, it has a coefficient of thermal expansion (CTE) of less than 2 × 10−10 /K [18]. To achieve laser frequency instability of 10−16 or even lower, careful considerations of passive thermal shielding and supporting configuration for reference cavities should be made. For systems that can not be temperature stabilized close to TCTE=0 or that have a lower temperature stability, it is also possible to realize high length stability of reference cavities by carefully designing their thermal shields [23].

This paper gives two models on how reference cavities respond to (1) a step environmental temperature change and (2) a periodic temperature fluctuation through heat transfer of thermal conduction and thermal radiation separately. The analysis shows that passive thermal shielding configurations of reference cavities with a larger thermal time constant will be less sensitive to environmental temperature fluctuations. All the thermal shields discussed in this paper are passive ones without active temperature control. With additional numerical simulation results based on finite element analysis, we will discuss the details on the design of cavity thermal shielding configurations to make it insensitive to environmental temperature perturbations, including material, layer numbers and aperture size of thermal shields. The number of thermal shields has a great effect on the time constant of reference cavities. A two-layer thermal shield will help to reduce the temperature fluctuation-induced length instability of a reference cavity below 1×10−15 on a day timescale when the cavity has a CTE of 1×10−10 /K and is enclosed in a vacuum chamber with temperature fluctuation amplitude of 1 mK and period of 24 hours.

2. Time constant

2.1. Thermal conduction

Consider body M1 that has a length of L1, cross-section area of A1 and heat conductivity of k1, as shown in Fig. 1(a). Another body M2 with a mass of m2 and heat capacity of C2 is on the top of M1. Assume that M2 is in good thermal contact with M1, and both are initially in thermal equilibrium.

 figure: Fig. 1

Fig. 1 Thermal conduction. (a) One-layer thermal conduction. (b) Two-layer thermal conduction. (c) The temperature variation of different layers over time when the temperature of M1 changes from Ti = 300 K to Tf = 350 K at t = 0.

Download Full Size | PDF

At time t = 0, the temperature at the bottom of M1 changes from Ti to TfT1 = TfTi). If ignoring the heat absorbed by M1 itself when M1 has a much smaller mass and thermal capacity compared with those of M2, the heat flow rate transferred between M1 and M2 is given by

dQdt=k1A1L1(TfT2),
where Q is transferred heat and T2 is the temperature of M2. M2 absorbs the heat that flows from M1, resulting in a temperature change
dT2=dQC2m2.
By solving the above equations, we obtain the temperature change of M2 over time
T2(t)=TfΔT1·ek1A1C2m2L1t=TfΔT1·etτc1.
It takes M2 time of τc1=C2m2L1k1A1 (time constant of thermal conduction) to change its temperature by (11e)ΔT1.

For the case of thermal conduction in a typical supporting configuration of reference cavities, M1 could be supporting rods of reference cavities or supporting rods of thermal shields. To enlarge the time constant, those supporting rods might have small thermal conductivity and cross-section area, while thermal shields on the top of the supporting rods have a large thermal capacity and mass. The red dashed line in Fig. 1(c) shows an illustration of the temperature change of M2 (made of copper with m2 = 20 kg) assuming the temperature of M1 (Teflon cylinders with L1 = 3 cm and total cross-section area A1 = 10−3 m2) changes from 300 K to 350 K. Here the time constant of thermal conduction is τc1 ≈ 256.7 h. Parameters for materials used throughout this paper can be found in Table 1.

Tables Icon

Table 1. Parameters of materials.

Suppose there are other layers M3 and M4 on the top of M2, as shown in Fig. 1(b). The temperature change of M4 is determined by

dT4dt=k3A3C4m4L3(TfΔT1etτc1T4).
Thus the temperature change of M4 which responds to the temperature variation of M1 is
T4(t)=TfΔT1τc1τc2(τc2etτc1τc2etτc2),
where τc1=C2m2L1k1A1 and τc2=C4m4L3k3A3. If M3 is also Teflon rods with L3 = 3 cm and A3 = 10−3 m2, and M4 is a piece of thermal shield made of copper with m4 = 15 kg, the time constants are τc1 ≈ 256.7 h and τc2 ≈ 192.5 h. The blue short dashed line in Fig. 1(c) shows how M4 responds to the step temperature change of M1. It takes M4 about 480.0 h to change its temperature by (11e)ΔT1. As it shows, in the two-layer thermal conduction the time constant is almost twice of that in the single-layer thermal conduction.

2.2. Thermal radiation

Thermal radiation is a dominant way of heat transfer for optical reference cavities since those cavities are usually put in vacuum chambers. Suppose there is an outer layer P1 and an inner layer P2, as shown in Fig. 2(a). Layer P1 could be a vacuum chamber, and P2 could be a layer of thermal shield. Initially, both of them are in thermal equilibrium. At t = 0, the temperature of P1 changes from Ti to TfT1 = TfTi). According to ref. [25], when ignoring the geometry of radiation bodies and distances between radiation surfaces, and supposing the view factor of radiation surfaces is unity (Indeed, the value of view factor is less than 1, which is related to radiation angle and relative spacing of radiation surfaces.), the rate of radiated energy absorbed by P2 is

dQdt=σ(Tf4T24)1A1(1ε11)+1A21ε2,
where A1 and A2 are inner surface area of P1 and outer surface area of P2 respectively, and σ = 5.67 × 10−8 W/(m2·K4) is the Stefan-Boltzmann constant. ε1 and ε2 are the emissivity of the corresponding surfaces of P1 and P2 respectively. Since the emissivity of a material varies as a function of temperature and surface finish, the exact emissivity of a material should be determined with absolute measurements. All the emissivity values concerning in this paper refer to Table 1, which are used just for reference. The temperature of P2 changes by
dT2dt=σ(Tf4T24)C2m2[1A1(1ε11)+1A21ε2]=β12(Tf4T24).
Using the first order approximation when ΔT1 is much smaller than Tf and Ti, the temperature variation of P2 over time is given by
T2(t)TfΔT1e4β12Tf3t=TfΔT1etτr1,
where τr1=14β12Tf3 is the time constant of thermal radiation for P2 to respond to the temperature change of P1. Note that in Eq. (8) τr1 depends on Tf instead of Ti due to the approximation(ΔT1Ti, Tf), while in fact τr1 has a weak dependence on Ti. This approximation gives an error of less than 5% for τr1 when ΔT1 < 10 K, Tf and Ti ≈ 300 K. The red dashed line in Fig. 2(c) shows an illustration of the temperature change of P2 over time when the temperature of P1 changes from 300 K to 310 K, assuming P1 is made of aluminum with ε1 = 0.2 and A1 = 1 m2, and P2 is made of copper with m2 = 20 kg, A2 = 0.6 m2 and ε2 = 0.1. The time constant is τr1 ≈ 6.5 h.

 figure: Fig. 2

Fig. 2 Thermal radiation. (a) One-layer thermal radiation. (b) Two-layer thermal radiation. (c) The temperature change of different layers over time when the temperature of P1 changes from Ti = 300 K to Tf = 310 K at t = 0.

Download Full Size | PDF

If there is another layer P3, which could be a cavity, inside P2, as shown in Fig. 2(b), the temperature change of P3 is determined by

dT3dt=β23[(TfΔT1etτr1)4T34],
where β23=σC3m3[1A2(1ε21)+1A31ε3]. By solving Eq. (9) to the first order approximation, we have the temperature change of P3, which is
T3(t)TfΔT1τr1τr2[τr1etτr1τr2etτr2],
where τr1=14β12Tf3 and τr2=14β23Tf3. If P3 is made of ULE glass with A3 = 0.2 m2, m3 = 10 kg and ε3 = 0.85, τr2 ≈ 6.6 h. The blue short dashed line in Fig. 2(c) shows the response of P3 to the step temperature change of P1 of 10 K. It takes P3 about 14.2 h to change its temperature by (11e)ΔT1. The green dash-dot line shows the response of P3 to the step temperature change of P1 if the intermediate layer P2 is missing (τr ≈ 3.1 h). As shown, the intermediate layer P2 helps to enlarge the time constant of P3 by more than four times.

3. Thermal sensitivity

Next we will discuss the situation in which the environmental temperature fluctuates periodically, we will see how reference cavities respond to the temperature variations.

3.1. Thermal conduction

In Fig. 1(a), suppose the temperature at the bottom of M1 fluctuates as T1(t)=T¯+ΔT1sin(2πtξ), in which is the mean temperature, ΔT1 and ξ are the amplitude and period of the temperature fluctuation respectively. Then the temperature of M2 changes periodically due to the existence of the heat transfer between M1 and M2 via thermal conduction. We use similar equations as Eqs. (1) and (2), and obtain the temperature of M2 as

dT2dt=1τc1[T¯+ΔT1sin(2πtξ)T2].
By solving the above equation, we have
T2(t)=T¯+ξξ2+4π2τc12ΔT1sin(2πtξϕ)+2πξτc1ξ2+4π2τc12ΔT1etτc1.
Figure 3(a) shows the temperature change of M2 when the temperature at the bottom of M1 fluctuates as T1(t)=300+10sin(2πt5×105). The large fluctuation period (ξ = 5 × 105 s) used here is to make the temperature fluctuation of M2 large enough to be clearly seen in the figure. As shown, the temperature of M2 fluctuates with the same period as that of M1, but it has a phase lag ϕ=arcsin(2πτc1ξ2+4π2τc12). And the mean value of T2(t) does not equal to due to the last term of Eq. (12). The temperature fluctuation amplitude of M2 is ΔT2=ξΔT1ξ2+4π2τc12.

 figure: Fig. 3

Fig. 3 Thermal sensitivity. (a) The temperature change of M2 when the temperature of M1 fluctuates as T1(t)=300+10sin[2πt5×105]. (b) The sensitivity of thermal conduction as a function of temperature fluctuation period. S (S2) is the sensitivity of one (two)-layer thermal conduction. (c) The sensitivity of thermal radiation as a function of temperature fluctuation period. S (S2) is the sensitivity of one (two)-layer thermal radiation.

Download Full Size | PDF

The normalized sensitivity of M2 to the temperature fluctuation of M1 can be written

S=ΔT2ΔT1=ξξ2+4π2τc12.
As shown in Eq. (13), if the period of the temperature fluctuation is much smaller than the time constant (ξτc1), Sξ2πτc1. In the other limit, if ξτc1, S ≈ 1.

If there are more layers on the top of M2, then the thermal sensitivity of Mn is

Sn=i=1nξξ2+4π2τci2.
Figure 3(b) shows the thermal sensitivity of conduction when using τc1 ≈ 256.7 h for S (the blue solid line) and τc1 ≈ 256.7 h and τc2 ≈ 192.5 h for two-layer sensitivity S2 (the red dashed line). As we can see in the figure, multiple-layer structure helps to make reference cavities less sensitive to temperature fluctuations with short fluctuation periods. For example, as seen in Fig. 3(b), when the temperature perturbations have fluctuation periods of 1 s to 106 s, the two-layer thermal sensitivity is S2S2.

3.2. Thermal radiation

For the model shown in Fig. 2(a), if the temperature of P1 fluctuates as T1(t)=T¯+ΔT1sin(2πtξ), due to thermal radiation the temperature of the inner layer P2 varies as

dT2dt=β12[(T¯+ΔT1sin(2πtξ))4T24].
When ΔT1 is small, the temperature variation of P2 is obtained to the first order approximation as
T2(t)T¯+ξξ2+4π2τr12ΔT1sin(2πtξψ)+2πξτr1ξ2+4π2τr12ΔT1etτr1,
where ψ=arcsin(2πτr1ξ2+4π2τr12). As we can see from the above equation, the temperature of P2 also fluctuates with the same period as that of P1. And the normalized sensitivity of P2 to the temperature fluctuation of P1 is given by
S=ξξ2+4π2τr12.
When the outer temperature fluctuation period ξ is much smaller compared to the thermal time constant of radiation τr1, Sξ2πτr1. In the opposite limit, if ξτr1, S ≈ 1.

If there are other layers inside P2, the thermal sensitivity of the most inside layer Pn to the temperature fluctuation of P1 is given by

Sn=i=1nξξ2+4π2τri2.
Figure 3(c) shows the thermal sensitivity of P2 when using τr1 ≈ 6.5 h (the blue solid line) and the thermal sensitivity of P3 with P2 as an intermediate layer using τr1 ≈ 6.5 h and τr2 ≈ 6.6 h (the red dashed line). The green dash-dot line is the thermal sensitivity of P3 without P2 using τr1 ≈ 3.1 h. As we can see that multiple-layer thermal radiation also helps to make reference cavities less sensitive to temperature fluctuations with short fluctuation period.

4. Simulation results for less thermal sensitivity

In the above analysis, whenever heat transfers via thermal conduction or thermal radiation, cavity thermal shielding configurations with a large time constant are insensitive to temperature fluctuations. Thereby, it is significant to enlarge the time constant of reference cavities.

Considering thermal conduction in an apparatus of a reference cavity, its thermal time constant (τc) is proportional to thermal capacity and mass of thermal shields and reference cavities, and inversely proportional to thermal conductivity and cross section area of supporting rods. To enlarge τc, material with small thermal conductivity should be chosen for supporting rods, such as vacuum compatible ceramic or Teflon. The geometry of those supporting rods is usually designed to be small in radius and large in length as long as it is stable enough to support.

For thermal radiation in an apparatus of a reference cavity, its thermal time constant τr is also proportional to thermal capacity and mass of thermal shields and reference cavities. For those reasons, the thermal shields for reference cavities are usually have a large mass. In the case of limited weight, aluminum is a good choice since its density is nearly one-third of that of copper while its heat capacity is three times bigger if without considering emissivity. To reduce thermal radiation between thermal shields and reference cavities, the emissivity of thermal shields should be small. Even for a specified material, the emissivity varies according to its surface treatment. Usually highly polished surfaces have lower emissivity, though it is technically difficult [24].

From the estimations of time constant listed above, the time constant of thermal radiation is more than an order of magnitude smaller than that of thermal conduction. Therefore thermal conduction could be neglected when calculating the combined time constant of reference cavities. It is also true when reference cavities are designed to be even longer in length for lower thermal noise [18, 19, 26], which makes the geometry of thermal shields even bigger. Since the mass as well as the radiation areas of thermal shields increase, the time constant of thermal conduction increases more than that of thermal radiation.

In the above analysis, both time constant and thermal sensitivity of thermal radiation are given to the first order approximation. Moreover, the geometry of radiation bodies and the distance between radiation surfaces are not taken into consideration, which affect the view factor of radiation and thus time constant of reference cavities as well. Therefore, in the following section, finite element analysis is performed in order to obtain a clearer and more accurate answer on the design of cavity shielding configurations, aiming at making reference cavities less sensitive to environmental temperature perturbations.

4.1. Material for thermal shields

We use ANSYS software to numerically simulate the temperature change of a reference cavity over time. The simulation model is shown in Fig. 4. A reference cavity and thermal shields are in the shape of cylinder. The cavity is made of ULE glass, and the outer vacuum chamber is made of aluminum. Only one layer of thermal shield, layer B as shown in the figure, is used to study the effect of different materials of thermal shield on the thermal time constant. The initial temperature of the whole setup is 25 °C. At t = 0, the temperature of the outer vacuum chamber raises to 30 °C.

 figure: Fig. 4

Fig. 4 The simulation model. The outer layer is a vacuum chamber made of aluminum, while the most inner layer is a reference cavity made of ULE glass. There are three layers of thermal shields inserted between the vacuum chamber and the reference cavity, labeled as A, B and C. All are in the shape of cylinder.

Download Full Size | PDF

The simulation results are shown in Fig. 5(a) and listed in Table 2. The thermal time constant of the reference cavity is 26.3 h, 23.8 h and 11.0 h when the thermal shield with the same size is made of gold, copper (roughly polished) and aluminum (roughly polished), respectively. For aluminum thermal shield, if radiation surfaces are highly polished, the emissivity could be smaller, e.g. 0.07, resulting in a thermal time constant of 26.0 h for the reference cavity. For the copper thermal shield, if it is plated by gold, the time constant increases to 32.5 h.

Tables Icon

Table 2. The thermal time constants when the material of thermal shield varies.

When using Eq. (10), the calculated time constants are also listed in the table, which are slightly smaller than the results based on finite element analysis since finite element analysis takes the geometry of each part, radiation angle and relative spacing of radiation surfaces into consideration. While in the calculation the thermal radiation is more efficient since the view factor is assumed to be unity. However, the calculation gives an estimation.

If there are three supporting rods made of Teflon with a diameter of 20 mm and length of 50 mm inserted between layers, a similar simulation is performed when considering both thermal radiation and thermal conduction. The combined thermal time constant of the reference cavity is 31.9 h when the thermal shield is made of gold-plated copper, as shown in Fig. 5(a). Comparing with the time constant of 32.5 h in the case that without supporting rods, it indicates that thermal conductivity plays a less significant role.

 figure: Fig. 5

Fig. 5 The simulation results of temperature change for the reference cavity over time when the temperature of outer vacuum chamber is changed from 25 °C to 30 °C if (a) the material of one-layer thermal shield varies and (b) the layer number of thermal shields varies. (c) The simulated and calculated thermal sensitivity of a reference cavity to outer temperature fluctuation when it is enclosed in zero to two layers thermal shields. (d) The simulated time constant of a reference cavity when the aperture size of the three-layer thermal shield varies.

Download Full Size | PDF

4.2. Layers of thermal shield

To study the effect of multiple-layer thermal shields, we numerically simulated the temperature variation of the reference cavity when the temperature of the outer vacuum chamber changes from 25 °C to 30 °C using the same model with different layer numbers of thermal shield made of gold-plated copper. As shown in Fig. 5(b), the time constant of the reference cavity is 1.5 h, 32.5 h, 87.2 h and 153.8 h if there are zero to three layers of thermal shields, respectively. This indicates that when adding one more layer, the thermal time constant of the reference cavity increases significantly.

The thermal sensitivities of the reference cavity as a function of temperature fluctuation period are also simulated when it is enclosed in different layers of thermal shields, as shown in Fig. 5(c). For each data in the figure, the temperature of the outer vacuum chamber fluctuates at a certain period with different amplitude ΔTout. By simulation, the temperature change of the inner reference cavity ΔTcav is obtained separately and thus the thermal sensitivity S = ΔTcavTout is obtained. In the figure, the simulation results of the sensitivities for reference cavities with no thermal shield, one-layer thermal shield and two-layer thermal shield are shown with dots, squares and triangles respectively, while the corresponding calculations results based on Eq. (10) are shown with dashed lines, which agree with the simulation results quite well.

From the above analysis, we can see that when using multiple-layer thermal shield the thermal time constant of reference cavities gets larger and reference cavities become less sensitive to environmental temperature fluctuations. However, it is not necessary to use multiple-layer thermal shields in each case. In Table 3, the temperature fluctuation-induced fractional length instability of reference cavities (ΔL/L) are estimated according to achievable outer temperature instability and the CTE of reference cavities. For example, if an outer vacuum chamber is temperature-stabilized with instability within 1 mK (24 hour fluctuation period) near TCTE=0, which might have a CTE of 1× 10−10/K, then the reference cavity can have a length instability at the 10−16 level on hundreds of seconds’ timescale even without any thermal shield, assuming the thermal noise limit of the reference cavity is below 1 × 10−16. If a reference cavity can not be conveniently controlled near TCTE=0, for example, which has a CTE of 5 × 10−9/K (for ULE glass, about 30 °C above TCTE=0), in order to achieve 10−16 length instability in an averaging time above 100 s, it should be enclosed in thermal shields. If the temperature variation rate of the outer vacuum chamber is large, the need for multiple-layer thermal shield becomes significant. Moreover, multiple-layer thermal shield is also important to achieve a high frequency stability on the timescale over a day. For example, as shown in Table 3, if the environmental temperature fluctuates with amplitude of 1 mK and period of 24 hours, a reference cavity with a CTE of 1 × 10−10/K is enclosed in a two-layer thermal shield, the temperature-induced length instability of the reference cavity can be at the 10−16 level over one day.

Tables Icon

Table 3. Estimation of temperature fluctuation-induced fractional length instability based on outer temperature variation rate, the CTE and the sensitivity of a reference cavity.

4.3. Aperture size of thermal shield

In the previous simulations, the thermal shields do not have any holes to let laser light access the reference cavities. However, all designs are going to need an aperture. To study the effect of the aperture size in the thermal shield on the thermal time constant, we numerically analyze the thermal time constants of reference cavities when the thermal shields having optical apertures with different diameters. In the case of one-layer thermal shield, layer B, the time constant decreases from 31.6 h to 30.0 h (changes by 5%) if the diameter of aperture in the thermal shield increases from 5 mm to 20 mm. If the reference cavity is enclosed in three-layer gold-plated copper thermal shields that have optical apertures d of 5 mm to 20 mm, the time constant decreases from 151 h to 135 h (changes by 12%), as shown in Fig. 5(d). This implies that the aperture size of thermal shields has a relatively weak effect on the thermal time constant compared to the material and layer numbers of thermal shields.

5. Conclusion

In this paper, thermal time constant and temperature sensitivity of reference cavities are studied in order to describe how reference cavities respond to environmental temperature fluctuations via heat transfer of thermal conduction and thermal radiation. The thermal sensitivity of reference cavities is proportional to fluctuation periods of temperature perturbations, and inversely proportional to the time constant of reference cavities. Both calculations based on equations and numerical simulations indicate that thermal radiation plays a significant role in enlarging the thermal time constant of reference cavities. To make reference cavities less sensitive to temperature perturbations, they might be enclosed in multiple layers of thermal shields with large mass, high thermal capacity and small emissivity. One choice for the material of thermal shields is copper with low emissivity. The aperture size of thermal shields has a weak effect on the thermal time constant of reference cavities. The layer number of thermal shields can be chosen according to experimentally achievable temperature stability and the CTE of reference cavities. By using multiple-layer thermal shield, the temperature fluctuation-induced fractional length instability at the 10−16 level on a day timescale could be achieved.

Acknowledgments

The authors would like to thank Chad Hoyt and Wei Zhang for helpful discussions. This work was supported by the National Natural Science Foundation of China (grant Nos. 11374102, 11104077, 11334002, 11127405 and 11475063) and the National Basic Research Program of China (grant No. 2012CB821302).

References and links

1. J. L. Hall, “Nobel lecture: defining and measuring optical frequencies,” Rev. Mod. Phys. 78(4), 1279–1295 (2006). [CrossRef]  

2. T. W. Hansch, “Nobel lecture: passion for precision,” Rev. Mod. Phys. 78(4), 1297–1309 (2006). [CrossRef]  

3. N. Hinkley, J. A. Sherman, N. B. Phillips, M. Schioppo, N. D. Lemke, K. Beloy, M. Pizzocaro, C. W. Oates, and A. D. Ludlow, “An atomic clock with 10−18 instability,” Science 341(6151), 1215–1218 (2013). [CrossRef]   [PubMed]  

4. B. J. Bloom, T. L. Nicholson, J. R. Williams, S. L. Campbell, M. Bishof, X. Zhang, W. Zhang, S. L. Bromley, and J. Ye, “An optical lattice clock with accuracy and stability at the 10−18 level,” Nature 506(7486), 71–75 (2014). [CrossRef]   [PubMed]  

5. T. M. Fortier, M. S. Kirchner, F. Quinlan, J. Taylor, J. C. Bergquist, T. Rosenband, N. Lemke, A. Ludlow, Y. Jiang, C. W. Oates, and S. A. Diddams, “Generation of ultrastable microwaves via optical frequency division,” Nat. Photonics 5(7), 425–429 (2011). [CrossRef]  

6. T. M. Fortier, N. Ashby, J. C. Bergquist, M. J. Delaney, S. A. Diddams, T. P. Heavner, L. Hollberg, W. M. Itano, S. R. Jefferts, K. Kim, F. Levi, L. Lorini, W. H. Oskay, T. E. Parker, J. Shirley, and J. E. Stalnaker, “Precision atomic spectroscopy for improved limits on variation of the fine structure constant and local position invariance,” Phys. Rev. Lett. 98(7), 070801 (2007). [CrossRef]   [PubMed]  

7. S. G. Turyshev, “Experimental tests of general relativity: recent progress and future directions,” Phys. Uspekhi 52(1), 1–27 (2009). [CrossRef]  

8. 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(2), 97–105 (1983). [CrossRef]  

9. A. D. Ludlow, X. Huang, M. Notcutt, T. Zanon-Willette, S. M. Foreman, M. M. Boyd, S. Blatt, and J. Ye, “Compact, thermal-noise-limited optical cavity for diode laser stabilization at 1 × 10−15,” Opt. Lett. 32(6), 641–643 (2007). [CrossRef]   [PubMed]  

10. H. Chen, Y. Jiang, S. Fang, Z. Bi, and L. Ma, “Frequency stabilization of Nd:YAG lasers with a most probable linewidth of 0.6 Hz,” J. Opt. Soc. Am. B 30(6), 1546–1550 (2013). [CrossRef]  

11. B. C. Young, F. C. Cruz, W. M. Itano, and J. C. Bergquist, “Visible lasers with subhertz linewidths,” Phys. Rev. Lett. 82(19), 3799–3802 (1999). [CrossRef]  

12. 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,” Nat. Photonics 6(10), 687–692 (2012). [CrossRef]  

13. L. Chen, J. L. Hall, J. Ye, T. Yang, E. Zang, and T. Li, “Vibration-induced elastic deformation of Fabry-Perot cavities,” Phys. Rev. A 74(5), 053801 (2006). [CrossRef]  

14. H. Stoehr, F. Mensing, J. Helmcke, and U. Sterr, “Diode laser with 1 Hz linewidth,” Opt. Lett. 31(6), 736–738 (2006). [CrossRef]   [PubMed]  

15. D. R. Leibrandt, M. J. Thorpe, M. Notcutt, R. E. Drullinger, T. Rosenband, and J. C. Bergquist, “Spherical reference cavities for frequency stabilization of lasers in non-laboratory environments,” Opt. Express 19(4), 3471–3482 (2011). [CrossRef]   [PubMed]  

16. Y. N. Zhao, J. Zhang, A. Stejskal, T. Liu, V. Elman, Z. H. Lu, and L. J. Wang, “A vibration-insensitive optical cavity and absolute determination of its ultrahigh stability,” Opt. Express 17(11), 8970–8982 (2009). [CrossRef]   [PubMed]  

17. B. Argence, E. Prevost, T. Leveque, R. Le Goff, S. Bize, P. Lemonde, and G. Santarelli, “Prototype of an ultra-stable optical cavity for space applications,” Opt. Express 20(23), 25409–25420 (2012) [CrossRef]   [PubMed]  

18. Y. Y. Jiang, A. D. Ludlow, N. D. Lemke, R. W. Fox, J. A. Sherman, L. S. Ma, and C. W. Oates, “Making optical atomic clocks more stable with 10−16-level laser stabilization,” Nat. Photonics 5(3), 158–161 (2011). [CrossRef]  

19. T. L. Nicholson, M. J. Martin, J. R. Williams, B. J. Bloom, M. Bishof, M. D. Swallows, S. L. Campbell, and J. Ye, “Comparison of two independent Sr optical clocks with 1 × 10−17 stability at 103 s,” Phys. Rev. Lett. 109(23), 230801 (2012). [CrossRef]  

20. K. Numata, A. Kemery, and J. Camp, “Thermal-noise limit in the frequency stabilization of lasers with rigid cavities,” Phys. Rev. Lett. 93, 250602 (2004). [CrossRef]  

21. P. Dube, A. A. Madej, J. E. Bernard, L. Marmet, and A. D. Shiner, “A narrow linewidth and frequency-stable probe laser source for the 88Sr+ single ion optical frequency standard,” Appl. Phys. B 95(1), 43–54 (2009). [CrossRef]  

22. J. Alnis, A. Matveev, N. Kolachevsky, Th. Udem, and T. W. Hansch, “Subhertz linewidth diode lasers by stabilization to vibrationally and thermally compensated ultralow-expansion glass Fabry-Perot cavities,” Phys. Rev. A 77, 053809 (2008). [CrossRef]  

23. C. Hagemann, C. Grebing, C. Lisdat, S. Falke, T. Legero, U. Sterr, F. Riehle, M. J. Martin, and J. Ye, “Ultrastable laser with average fractional frequency drift rate below 5 × 10−19/s,” Opt. Lett. 39(17), 5102–5105 (2014). [CrossRef]   [PubMed]  

24. “Table of total emissivity,” www.omega.com/temperature/Z/pdf/z088-089.pdf.

25. J. G. Knudsen, H. C. Hottel, A. F. Sarofim, P. C. Wankat, and K. S. Knaebel, “Heat and mass transfer,” in Perry’s Chemical Engineers’ Handbook, R. H. Perry and D. W. Green, eds. 8th ed. (Academic, 2007), pp. 25–32.

26. M. Notcutt, L. S. Ma, A. D. Ludlow, S. M. Foreman, J. Ye, and J. L. Hall, “Contribution of thermal noise to frequency stability of rigid optical cavity via hertz-linewidth lasers,” Phys. Rev. A 73, 031804(R) (2006). [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 Thermal conduction. (a) One-layer thermal conduction. (b) Two-layer thermal conduction. (c) The temperature variation of different layers over time when the temperature of M1 changes from Ti = 300 K to Tf = 350 K at t = 0.
Fig. 2
Fig. 2 Thermal radiation. (a) One-layer thermal radiation. (b) Two-layer thermal radiation. (c) The temperature change of different layers over time when the temperature of P1 changes from Ti = 300 K to Tf = 310 K at t = 0.
Fig. 3
Fig. 3 Thermal sensitivity. (a) The temperature change of M2 when the temperature of M1 fluctuates as T 1 ( t ) = 300 + 10 sin [ 2 π t 5 × 10 5 ]. (b) The sensitivity of thermal conduction as a function of temperature fluctuation period. S (S2) is the sensitivity of one (two)-layer thermal conduction. (c) The sensitivity of thermal radiation as a function of temperature fluctuation period. S (S2) is the sensitivity of one (two)-layer thermal radiation.
Fig. 4
Fig. 4 The simulation model. The outer layer is a vacuum chamber made of aluminum, while the most inner layer is a reference cavity made of ULE glass. There are three layers of thermal shields inserted between the vacuum chamber and the reference cavity, labeled as A, B and C. All are in the shape of cylinder.
Fig. 5
Fig. 5 The simulation results of temperature change for the reference cavity over time when the temperature of outer vacuum chamber is changed from 25 °C to 30 °C if (a) the material of one-layer thermal shield varies and (b) the layer number of thermal shields varies. (c) The simulated and calculated thermal sensitivity of a reference cavity to outer temperature fluctuation when it is enclosed in zero to two layers thermal shields. (d) The simulated time constant of a reference cavity when the aperture size of the three-layer thermal shield varies.

Tables (3)

Tables Icon

Table 1 Parameters of materials.

Tables Icon

Table 2 The thermal time constants when the material of thermal shield varies.

Tables Icon

Table 3 Estimation of temperature fluctuation-induced fractional length instability based on outer temperature variation rate, the CTE and the sensitivity of a reference cavity.

Equations (18)

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

d Q d t = k 1 A 1 L 1 ( T f T 2 ) ,
d T 2 = d Q C 2 m 2 .
T 2 ( t ) = T f Δ T 1 · e k 1 A 1 C 2 m 2 L 1 t = T f Δ T 1 · e t τ c 1 .
d T 4 d t = k 3 A 3 C 4 m 4 L 3 ( T f Δ T 1 e t τ c 1 T 4 ) .
T 4 ( t ) = T f Δ T 1 τ c 1 τ c 2 ( τ c 2 e t τ c 1 τ c 2 e t τ c 2 ) ,
d Q d t = σ ( T f 4 T 2 4 ) 1 A 1 ( 1 ε 1 1 ) + 1 A 2 1 ε 2 ,
d T 2 d t = σ ( T f 4 T 2 4 ) C 2 m 2 [ 1 A 1 ( 1 ε 1 1 ) + 1 A 2 1 ε 2 ] = β 12 ( T f 4 T 2 4 ) .
T 2 ( t ) T f Δ T 1 e 4 β 12 T f 3 t = T f Δ T 1 e t τ r 1 ,
d T 3 d t = β 23 [ ( T f Δ T 1 e t τ r 1 ) 4 T 3 4 ] ,
T 3 ( t ) T f Δ T 1 τ r 1 τ r 2 [ τ r 1 e t τ r 1 τ r 2 e t τ r 2 ] ,
d T 2 d t = 1 τ c 1 [ T ¯ + Δ T 1 sin ( 2 π t ξ ) T 2 ] .
T 2 ( t ) = T ¯ + ξ ξ 2 + 4 π 2 τ c 1 2 Δ T 1 sin ( 2 π t ξ ϕ ) + 2 π ξ τ c 1 ξ 2 + 4 π 2 τ c 1 2 Δ T 1 e t τ c 1 .
S = Δ T 2 Δ T 1 = ξ ξ 2 + 4 π 2 τ c 1 2 .
S n = i = 1 n ξ ξ 2 + 4 π 2 τ c i 2 .
d T 2 d t = β 12 [ ( T ¯ + Δ T 1 sin ( 2 π t ξ ) ) 4 T 2 4 ] .
T 2 ( t ) T ¯ + ξ ξ 2 + 4 π 2 τ r 1 2 Δ T 1 sin ( 2 π t ξ ψ ) + 2 π ξ τ r 1 ξ 2 + 4 π 2 τ r 1 2 Δ T 1 e t τ r 1 ,
S = ξ ξ 2 + 4 π 2 τ r 1 2 .
S n = i = 1 n ξ ξ 2 + 4 π 2 τ r i 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.