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

Thermal lensing in optical fibers

Open Access Open Access

Abstract

Average powers from fiber lasers have reached the point that a quantitative understanding of thermal lensing and its impact on transverse mode instability is becoming critical. Although thermal lensing is well known qualitatively, there is a general lack of a simple method for quantitative analysis. In this work, we first conduct a study of thermal lensing in optical fibers based on a perturbation technique. The perturbation technique becomes increasingly inaccurate as thermal lensing gets stronger. It, however, provides a basis for determining a normalization factor to use in a more accurate numerical study. A simple thermal lensing threshold condition is developed. The impact of thermal lensing on transverse mode instability is also studied.

© 2016 Optical Society of America

Corrections

Liang Dong, "Thermal lensing in optical fibers: erratum," Opt. Express 31, 25042-25046 (2023)
https://opg.optica.org/oe/abstract.cfm?uri=oe-31-15-25042

1. Introduction

In the past decade, there has been significant progress in average power scaling of fiber lasers. Currently, fiber lasers are used in a wide range of industrial processes from marking and engraving to heavy duty cutting and welding. There is still a strong interest in further average power scaling of fiber lasers, mostly driven by new applications in science [1,2] and defense [3]. A quantitative understanding of thermal lensing is becoming increasingly critical for any further average power scaling efforts. It is also important to understand the impact of thermal lensing on transverse mode instability (TMI) [4–6].

Considering the importance of thermal lensing, there are surprisingly few published general studies of this effect in optical fibers. Dawson et al. studied the effect of thermal lensing based on a parabolic thermal profile and an ABCD matrix to approximate Gaussian beam propagation [7]. The guidance of the original optical fiber is ignored in this study. The balance of wave guidance due to thermal lensing and diffraction leads to an approximate power limit for thermal lensing. Such analysis is invalid at low heat loads where the original fiber design plays a dominant role in waveguiding. Thermal-optical effects were studied numerically for a certain fiber amplifier arrangement in [8]. Thermal lensing and transition of an original single-mode fiber to a multimode fiber at high heat load were observed. No systematic study was performed. Recently the reduction of mode field diameter (MFD) at high average powers was experimentally characterized for a number of optical fibers [9]. The observed MFD reduction was much more pronounced in fibers with larger MFD at a similar heat load. One of these fibers was simulated using a 3D model for its MFD reduction at high thermal loads [10]. Similar MFD reduction was also observed in thermally guiding and index-anti-guiding fiber [11]. A theoretical analysis of optical modes at various thermal loads for this type of fibers was also conducted [12]. No general studies of thermal lensing were performed in these works.

In this work, we attempt to develop a full quantitative analysis of thermal lensing in optical fibers. We first study thermal lensing in optical fibers based on a perturbation method. Although this analysis is inaccurate for strong thermal lensing, it does provide an insight into the key parameters involved in thermal lensing. We then use the normalized thermal lensing parameter developed in the perturbation study in a more accurate numerical model. A cylindrical layered optical mode solver is used in the numerical study. A numerical method for solving a heat conduction equation with a spatially non-uniform heat load and cylindrical symmetry is developed. The parabolic thermal profile based on approximated uniform heating which was used frequently in the past is found to underestimate the core temperature. A simple criteria for thermal lensing is developed. The impact of thermal lensing on TMI is also studied and TMI is found to be generally suppressed by thermal lensing. This is largely due to the fact that thermal lensing pulls power in optical modes towards the center of the optical fiber. This pulling effect is stronger for the fundamental mode than for higher-order modes, consequently reducing the overlap between the fundamental mode and higher-order modes.

2. Perturbation analysis of thermal lensing

The perturbation method used is very similar to that described in [13]. The thermal-optic effect is introduced by

n(r)=n0(r)+kTΔT(r)
where refractive index profile n, initial refractive index profile n0 and change in temperature ΔT are expressed in cylindrical coordinates, and kT = 1.1 × 10−5 K−1 is the thermal-optic coefficient for silica. Under the influence of heat load, the guided mode is expected shrink down to a new equilibrium where it would be stable. This is very similar to what happens under nonlinear self-focus [13]. The only difference is in the time taken to reach this new equilibrium. It is expected to be much slower and determined by the thermal diffusion rate perpendicular to the propagation direction. Under very strong thermal lensing, the optical mode can be focused down to a very small size over a distance in the order of the Rayleigh range, similar to that in nonlinear self-focus.

The temperature profile is approximated by the parabolic solution for uniform heating in the core with a heat load of Q0 (w/m) and we have ignored the temperature change outside the core by setting ΔT = 0 for r>ρ, where ρ is the core radius. This is reasonable for multimode fibers commonly found in fiber lasers as the optical power is mostly in the core in these cases. It is however not a good approximation for single-mode fibers where there is significant optical power outside the core.

ΔT(r)=Q04πκρ2(ρ2r2)
where κ = 1.38 w/m/K is the thermal conductivity. Details of the perturbation analysis are given in Appendix A. It is based on finding a relationship between modes in the fiber perturbed by the heat load and the original unperturbed fiber. With the assumption of a Gaussian mode (see Eq. (A11) in Appendix A), we can obtain an equation for the relative mode size γ = w/w0, where w is the spot size of the mode in the perturbed fiber and w0 in the unperturbed fiber. Spot size is half of the MFD for a Gaussian mode (see Eq. (A11) in Appendix A for a detailed definition).
12ξQ0w02γ6+(U2+3ξQ0ρ2+92ξQ0w02)γ4+3ξQ0ρ2γ2U2=0
where U is the core parameter as normally defined for optical fiber, and
ξ=kTk02nco24πκ
where k0 is the vacuum wave number and nco is the core refractive index. Equation (3) can be solved numerically and the result is shown in Fig. 1 for V = 3-8.

 figure: Fig. 1

Fig. 1 Mode collapse under the influence of thermal lensing. Equation (3) is evaluated for normalized frequency V = 3-8. A numerical simulation is also performed for comparison for fibers with core diameters of 10μm, 20μm and 30μm and NA of 0.06. The results are plotted as MFD for near-field MFD and eMFD for effective MFD, i.e. 2(Aeff/π) ½.

Download Full Size | PDF

The relative mode size is plotted against the normalized thermal lensing parameter ξQ0w02 in Fig. 1. It can be seen that the rate of change in the relative mode size is slow at small ξQ0w02, but accelerates at larger ξQ0w02. There is very little difference in the curves for different V values. The relative mode size can collapse to small values at high heat load and the rate of change slows down when ξQ0w02>2. The most interesting observation is the fact that the effect of thermal lensing is fully characterized by the normalized thermal lensing parameter ξQ0w02, where ξ is fully determined by material properties and laser wavelength. This normalized thermal lensing parameter scales linearly with heat load Q0 and also scales quadratically with MFD. This MFD dependence of thermal lensing was also shown in [7]. Since the total effect of thermal lensing is determined by the integrated effect seen by the entire mode, it is reasonable to expect it to scale with the mode area, i.e. w02.

3. Numerical analysis of thermal lensing

For a more accurate study of the impact of thermal lensing, we need to conduct a numerical study. We will focus on the regime where the fundamental mode dominates and ignore any mode distortion due to bending. Since we are only dealing with optical modes with a cylindrical symmetry in this case, we will use an optical mode solver for an arbitrary refractive index profile with a cylindrical symmetry. We already have established such a vector optical mode solver for an earlier work [14,15]. The formalism is very similar to that described in [16].

We also need to deal with the temperature profile resulting from the heat load from an optical mode with cylindrical symmetry. In our analysis, we have assumed that heat load is proportional to the local mode intensity; consequently we have ignored the effect of gain saturation. Gain saturation will lead to a flattening out of heat load in the center of the core and therefore slightly weaken the effect of thermal lensing.

Since a simple parabolic solution exists for the temperature profile in the case of a spatially uniform heat load in a cylinder, it has been used in many previous thermal analysis of fiber lasers [7,17]. We need to have a more accurate numerical model for the temperature profile in a cylinder given an arbitrary cylindrical heat load profile. We start by dividing the area with the heat source into many fine layers so that we can assume each layer has a uniform heat load and we can then use the parabolic solution over each layer. This approach, however, does not have enough free parameters to ensure the solution has both continuity and continuity in the 1st order derivative at the boundary between layers, both required for a rigorous solution. We subsequently devised a scheme where we first divide the area with the heat load into many fine layers with equal thickness. We then divide each layer into two sub-layers, one sub-layer with the total heat load for the layer and one sub-layer without source. We can then use the known parabolic solution over the sub-layer with the source and the known logarithmic solution for the sub-layer without source. The relative thickness of the sub-layers can be determined by the continuity condition at the boundary between layers. This new scheme allows us to find an accurate solution while ensuring necessary continuities at all the boundaries. A detailed derivation is provided in Appendix B. In the simulations in this work, the region with the heat load is typically divided into 100-200 layers. This is found to be sufficient for a stable solution.

A comparison of the temperature profiles from the numerical model used in this study and the uniform heat load is shown in Fig. 2. The fiber has a NA of 0.06 and a core diameter of 20μm, operated at 1.03μm. We use a cladding diameter of 400μm throughout this study. The total heat load in the core is 64w/m in both cases and there is no heat load outside the core. The normalized heat load profile, i.e. the normalized mode intensity profile, is also shown. The numerical model used an iteration process described in the next paragraph to find the optical mode under heat load. It can be seen clearly that the solution in the cladding is the same for both cases and the uniform heat load model underestimates the temperature in the core. This is due to the fact that more heat is deposited near the core center than that accounted in the uniform heat load model.

 figure: Fig. 2

Fig. 2 Comparison of the temperature profile from the numerical method used in this work and the parabolic solution for a uniform heat load. ΔT is set to be 0K at the cladding boundary. Core diameter is 20μm; cladding diameter is 400μm and NA is 0.06. The total heat load is 64w/m in both cases. The laser wavelength is at 1.03μm. The normalized heat load profile, i.e. normalized mode intensity profile, is also shown.

Download Full Size | PDF

In the following analysis of thermal lensing in an optical fiber, we first find the fundamental optical mode in the fiber without any heat load. We then apply one Mth of the heat load with a heat load profile equaling the fundamental optical mode intensity. The temperature profile is then calculated, which is then used to calculate the refractive index profile of the new fiber. The fundamental mode is then found for this new fiber. Two Mth of the heat load is then used with a heat load profile equaling the new fundamental mode intensity for the next iteration. This is repeated until total heat load is applied. This slow increase of heat load over many iterations is essential to keep the change in optical mode profile to a minimal for each iteration. Otherwise it can lead to numerical instabilities especially at large heat loads. We found M = 30 to be adequate to ensure stability and convergence for our analysis. This is used throughout this study.

The evolution of MFD (near-field MFD) and eMFD (effective MFD obtained from effective mode area) over the gradual increase of heat load for the case in Fig. 2 is shown in Fig. 3. The relative change in MFD and eMFD is shown on the right vertical axis. The reduction of mode size due to thermal lensing when heat load is gradually increased can be clearly seen. Also can be seen is that the mode size stabilizes at a constant value after the 31st run when the heat load is no longer increased. This demonstrates the convergence of the solution and stability of the numerical process.

 figure: Fig. 3

Fig. 3 Typical run for the case in Fig. 2. Core diameter is 20μm and NA is 0.06. The total heat load is 64w/m. The laser wavelength is at 1.03μm. Relative changes are shown on the vertical axis on the right. The heat load is gradually increased over 30 steps.

Download Full Size | PDF

The refractive index profiles obtained by the numerical model for the 20μm-core fiber are shown in Fig. 4 for various total heat loads. The refractive index profile is truncated at a radius 20μm, i.e. twice the core radius in this case. In our following analysis of optical mode, we are only concerned with the part of the waveguide seen by the optical mode of interest. We typically truncate the refractive index profile to 2 to 5 times of the core radius depending on guiding strength of the waveguide. Using an unnecessarily large cladding radius in the analysis can lead to poor numerical stability in the optical mode solver in addition to longer computation time. Care is taken in each case to ensure that the truncation does not compromise accuracy.

 figure: Fig. 4

Fig. 4 Refractive index profiles under various heat loads. Core diameter is 20μm and NA is 0.06. The laser wavelength is at 1.03μm.

Download Full Size | PDF

We then proceed to study the impact of thermal lensing on mode size for three fibers with core diameters of 10μm, 20μm and 30μm. All fibers have a NA of 0.06 and a cladding diameter of 400μm. This study is conducted for a wavelength of 1.06μm. The respective V values are 1.778, 3.557, and 5.335. The 10μm-core fiber is in the single-mode regime and the other two fibers are in the multimode regime. The results are shown in Fig. 1 for both MFD and eMFD (see figure caption for detailed definition). It can be seen that the results from the perturbation method are reasonable for a small normalized thermal lensing parameter, but overestimate the effect of thermal lensing when ξQ0w02>0.1. The results for the three fibers at ξQ0w02>2 are very close. In this regime, wave guidance is almost entirely from the effect of thermal lensing and the original waveguide plays a very small part. At smaller ξQ0w02, the single-mode fiber suffers slightly more mode size reduction. The mode is much larger than the core in the single-mode regime and this enhances the impact of thermal lensing on mode size. MFD and eMFD follows each other fairly closely. A convenient place to set the thermal lensing threshold is ξQ0w02 = 0.5, where the mode size is reduced by ~10% and the effective mode area by ~20%. The thermal lensing limit is put at ξQ0w02 = 0.18 in [7]. This study is able to quantify the change in mode size for this condition for the first time; the mode size is reduced by ~4% and the effective mode area by ~8% at ξQ0w02 = 0.18.

Using the numerical study shown in Fig. 1, we can also provide a theoretical mode size change for the experimental work in [9]. Quantum defect heating is used to convert heat load to extracted power. The average heat load over the amplifier is used, since this can be easily found in the paper but not the more relevant peak heat load. This will underestimate the effect of thermal lensing in this case. This is shown in Fig. 5. For three fibers with smallest MFDs, the predicted MFDs are very close to those measured (within few percent). The divergence becomes large for the fibers with large MFD (>80μm). The measured data for those fibers are highly scattered, indicating much larger measurement errors in this regime. This model does not consider effects such as anti-crossings with modes originated in the cladding and photo-darkening, which are known to take place in some of these fibers. It nevertheless provides a reasonable qualitative agreement with the measurements even for fibers with large MFD.

 figure: Fig. 5

Fig. 5 Simulated MFD for the fibers in [9]. Quantum defect heating is used to convert heat load to extracted power. Average heat load is used.

Download Full Size | PDF

4. The impact of thermal lensing on TMI

There is a strong interest in understanding the impact of thermal lensing on TMI. TMI can be quantified by the TMI nonlinear coefficient χ defined in [6]. TMI threshold is inverse proportional to χ and also dependent on input higher-order-mode power and amplifier configuration. Since we can easily evaluate the refractive index profile of the fiber under thermal loading, we just need to find the fundamental and higher-order modes to evaluate χ under the influence of thermal lensing. We have assumed that the fundamental mode is dominating in this case and conducted this study for the LP11 mode. Our cylindrical optical vector cannot directly find the LP11 mode, but can find its constituent TE01 and HE21 modes. These modes have the same radial intensity profile as LP11 (see Table14-1 in [18]). We therefore found the radial mode intensity profile for the HE21 mode and used this in Eq. (28) in [6] for our study.

We studied 5 fibers with core diameters of 10μm, 15μm, 20μm, 25μm and 30μm respectively. The fiber NA is 0.06 and cladding diameter is 400μm. This study is conducted at a wavelength 1.06μm. The V values are respectively 1.778, 2.667, 3.557, 4.446, and 5.335. The results are shown in Fig. 6. The 10μm-core fiber is single mode at low heat load, but the LP11 mode is guided when ξQ0w02>0.084. Its TMI nonlinear coupling coefficient χ increases until ξQ0w02 = ~1 as the LP11 mode is increasingly guided. When ξQ0w02>1, thermal lensing becomes significant and χ starts to decrease. Thermal lensing pulls all modes to the core center, but this is more significant for the fundamental mode (see Fig. 7). This lowers the overlap between the LP01 and LP11 modes, consequently leading to a reduction in χ. A similar trend can be seen for the 15μm-core fiber. Since the LP11 is already well guided in this fiber with heat load, the initial increase in χ at low heat load is less pronounced. χ starts to decrease when ξQ0w02>0.3. The remaining fibers have similar χ at low heat load; this is due to that fact that χ changes very slowly at large V [6]. For these fibers, χ starts to decrease significantly when ξQ0w02>0.3. Similar studies were conducted for LP02 mode and similar trends were obtained. This reduction of TMI at high thermal load is, however, of limited practical use as the effect of thermal lensing is significant at this point. For most practical fibers, the TMI threshold is also well below this thermal load.

 figure: Fig. 6

Fig. 6 Simulated TMI nonlinear coupling coefficient for fibers with core diameters of 10μm, 15μm, 20μm, 25μm and 30μm respectively. The fiber NA is 0.06 and cladding diameter is 400μm. This study is conducted at a wavelength of 1.06μm. Contour lines indicate constant thermal load.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 The LP01 and LP11 modes at 1.06μm in a fiber with a NA of 0.06 and a core diameter of 20μm without heat load and with a heat load of 1000w/m, i.e. ξQ0w02 = 2.19. The cladding diameter is 400μm.

Download Full Size | PDF

5. Conclusions

We have studied thermal lensing effects in optical fibers with both a perturbation method and numerical method and have developed a normalized thermal lensing parameter. A simple thermal lensing threshold condition is also developed. We have further studied the impact of thermal lensing on TMI and found that strong thermal lensing leads to a reduction in TMI.

6 Appendix A. Detailed perturbation analysis of thermal lensing

This derivation is very similar to that in [13]. Readers are encouraged to read [13] for a fuller description. With a perturbation of a temperature profile ΔT(r), the waveguide equation can be re-written in a normalized cylindrical coordination where R=wr and w is the spot size of the optical mode.

tψ(R)+{w2k02n0w2(R)[1+2kTΔT(R)]w2β2m2R2}ψ(R)=0
where ∇t = ∂2/∂R2 + (1/R)∂/∂R is the transverse operator; ψ(R) is radial field profile of the local mode; k0 is the vacuum wave number, m is the azimuthal mode number, β is the propagation constant of the local mode, and n0w is the index profile of the unperturbed fiber in the normalized coordination where R = wr.

For the unperturbed fiber, similarly we have

tψ0(R)+{w02k02n0w02(R)w02β02m2R2}ψ0(R)=0
where the subscript 0 indicates it belongs to mode of the unperturbed fiber, R = w0r is normalized against w0 in this case and n0w0 is the index profile of the unperturbed fiber in the normalized coordination where R = w0r.

We can now multiply Eq. 5 by ψ0 and then subtract Eq. 6 multiplied by ψ. The resulting equation is integrated over the cross section of the fiber, i.e. A∞. We then have

A[w02(k02n0w02β02)]w2(k02n0w2β2)ψψ0dA=2kTw2k02An0w2ΔTdA+Aψ0tψψtψ0dA
The last term in Eq. (7) can be converted into a line integral at infinity and is equal to zero for a confined optical mode. We then have
w2β2w02β02=2kTw2k02nT2k02(w02na2w2nb2)
This can be written as
w2=w02(k02na2β02)2kTw2k02nT2k02nb2β2
where

na2=An0w02(R)ψ(R)ψ0(R)dAAψ(R)ψ0(R)dA  nb2=An0w2(R)ψ(R)ψ0(R)dAAψ(R)ψ0(R)dA  nT2=An0w2(R)ΔT(R)ψ(R)ψ0(R)dAAψ(R)ψ0(R)dA

We then repeat this process with the waveguide equation in cylindrical coordination without normalization. The waveguide equation for the fiber perturbed by the temperature profile in this case is

tψ(r)+{k02n02(r)[1+2kTΔT(r)]β2m2r2}ψ(R)=0
The waveguide equation for the unperturbed fiber is
tψ0(r)+{k02n02(r)β02m2R2}ψ0(r)=0
Following a similar procedure to that for Eqs. (5) and 6, we have
β2β02=2kTk02nTr2
where

nTr2=An02(r)ΔT(r)ψ(r)ψ0(r)dAAψ(r)ψ0(r)dA

We can evaluate Eqs. 10 and 14 for the parabolic temperature expressed in Eq. 2 and Gaussian mode field profile expressed in

ψ(r)=e12r2w2
The results are

na2nco2nb2nco2nT2nco2Q08πκρ2(ρ2+w22)
nTr2nco2Q04πκρ2(ρ2+21w2+1w02)

Substituting Eq. 17 into Eq. 13, we have

β2β02=kTk02nco2Q02πκρ2(ρ2+21w2+1w02)

Substituting Eqs. 16 and 18 into Eq. 9, we have

12ξQ0w02γ6+(U2+3ξQ0ρ2+92ξQ0w02)γ4+3ξQ0ρ2γ2U2=0

where γ=w/w0, U is the core parameter as normally defined for optical fibers, and

ξ=kTk02nco24πκ

7 Appendix B. Temperature profile with arbitrary cylindrical heating

The cylindrical heat load is described by Q(r) in w/m/μm2 in a cylindrical coordination. The cladding radius is b. We will also assume that the fiber is cooled and the cladding surface temperature is constant at T0, i.e. T(b)= T0. The heat conduction equation in this case is given by

2T(r)r2+1rT(r)r+Q(r)κ=0
In areas with no source, the solution is known,
T(r)=C1+C2ln(r)
where C1 and C2 are constants to be determined by the boundary conditions. In areas with a uniform heat source with a heat load Q, the solution is also known
T(r)=CQ4κr2
where C is a constant to be determined by the boundary condition. The heat load extends to r = a. There is no heat load beyond r = a.

We first divide the area between r=0 and r=a into N concentric layers with an equal thickness of a/N. We further divide each layer into two concentric sub-layers, one with a constant source of the heat load for the layer with an outer edge at rna and one without any source with an outer edge at rn=n×a/N (see Fig. 8). Starting from the innermost circle, i.e. n=1, the solution for both the areas with source and without source can be found while ensuring continuity and the 1st order derivative continuity at the internal boundary. They are respectively

T1(r)=TcQ14κr12r1a2r2
T1a(r)=TcQ14κr12Q12κr12ln(rr1a)
where Tc is the temperature at the core center. For the next section n = 2, ensuring continuity at the boundary between the 1st and the 2nd layers and continuity and the 1st order derivative continuity at the internal boundary, we similarly have

 figure: Fig. 8

Fig. 8 Illustration of the scheme for solving the heat conduction equation with an arbitrary heat load profile with cylindrical symmetry.

Download Full Size | PDF

T2(r)=T1a(r1)+Q24κr22r12r2a2r12(r12r2)
T2a(r)=T1a(r1)+Q24κ(r12r22)Q22κr22r12r2a2r12r2a2ln(rr2a)

The continuity of the 1st order derivative at the boundary between the first and the second layer requires (note that r1a can be chosen arbitrarily)

r2a2=r12+Q2Q1(r22r12)

For the next layer n=3, we have

T3(r)=T2a(r2)+Q34κr32r22r3a2r22(r22r2)
T3a(r)=T2a(r2)+Q34κ(r22r32)Q32κr32r22r3a2r22r3a2ln(rr3a)

The continuity of the 1st order derivative at the boundary between the second and the third layer requires

r3a2=r22+Q3Q2(r32r22)r2a2r12r22r12r22r2a2

For n=N, we have

TN(r)=T(N1)a(rN1)+QN4κrN2rN12rNa2rN12(rN12r2)
TNa(r)=T(N1)a(rN1)+QN4κ(rN12rN2)QN2κrN2rN12rNa2rN12rNa2ln(rrNa)

The continuity of the 1st order derivative at the boundary between the (N-1)th and the Nth layer requires

rNa2=rN12+QNQN1(rN2rN12)r(N1)a2rN22rN12rN22rN12r(N1)a2

For area without source, i.e. r>a

T(r)=T(N1)a(rN1)+QN4κ(rN12rN2)QN2κrN2rN12rNa2rN12rNa2ln(rrNa)

The boundary condition T(b)=T0 requires

T0=T(N1)a(rN1)+QN4κ(rN12rN2)QN2κrN2rN12rNa2rN12rNa2ln(brNa)

Funding

Joint Technology Office High Energy Lasers MRI program (911NF-12-1-0332).

Acknowledgment

This material is based upon work supported in part by the U. S. Army Research Laboratory and the U. S. Army Research Office under contract/grant number W911NF-12-1-0332 through a Joint Technology Office MRI program.

References and links

1. T. Katsouleas, “Accelerator physics: Electrons hang ten on laser wake,” Nature 431(7008), 515–516 (2004). [CrossRef]   [PubMed]  

2. J. J. Macklin, J. D. Kmetec, and C. L. Gordon 3rd, “High-Order Harmonic Generation Using Intense Femtosecond Pulses,” Phys. Rev. Lett. 70(6), 766–769 (1993). [CrossRef]   [PubMed]  

3. J. Hecht, “Half a century of laser weapons,” Opt. Photonics News 20(2), 14–21 (2009). [CrossRef]  

4. T. Eidam, C. Wirth, C. Jauregui, F. Stutzki, F. Jansen, H. J. Otto, O. Schmidt, T. Schreiber, J. Limpert, and A. Tünnermann, “Experimental observations of the threshold-like onset of mode instabilities in high power fiber amplifiers,” Opt. Express 19(14), 13218–13224 (2011). [CrossRef]   [PubMed]  

5. A. V. Smith and J. J. Smith, “Mode instability in high power fiber amplifiers,” Opt. Express 19(11), 10180–10192 (2011). [CrossRef]   [PubMed]  

6. L. Dong, “Stimulated thermal Rayleigh scattering in optical fibers,” Opt. Express 21(3), 2642–2656 (2013). [CrossRef]   [PubMed]  

7. J. W. Dawson, M. J. Messerly, R. J. Beach, M. Y. Shverdin, E. A. Stappaerts, A. K. Sridharan, P. H. Pax, J. E. Heebner, C. W. Siders, and C. P. J. Barty, “Analysis of the scalability of diffraction-limited fiber lasers and amplifiers to high average power,” Opt. Express 16(17), 13240–13266 (2008). [CrossRef]   [PubMed]  

8. K. R. Hansen, T. T. Alkeskjold, J. Broeng, and J. Lægsgaard, “Thermo-optical effects in high-power ytterbium-doped fiber amplifiers,” Opt. Express 19(24), 23965–23980 (2011). [CrossRef]   [PubMed]  

9. F. Jansen, F. Stutzki, H. J. Otto, T. Eidam, A. Liem, C. Jauregui, J. Limpert, and A. Tünnermann, “Thermally induced waveguide changes in active fibers,” Opt. Express 20(4), 3997–4008 (2012). [CrossRef]   [PubMed]  

10. C. Jauregui, H. J. Otto, F. Stutzki, J. Limpert, and A. Tünnermann, “Simplified modelling the mode instability threshold of high power fiber amplifiers in the presence of photodarkening,” Opt. Express 23(16), 20203–20218 (2015). [CrossRef]   [PubMed]  

11. F. Jansen, F. Stutzki, H. J. Otto, C. Jauregui, J. Limpert, and A. Tünnermann, “High-power thermally guiding index-antiguiding-core fibers,” Opt. Lett. 38(4), 510–512 (2013). [CrossRef]   [PubMed]  

12. L. Kong, J. Cao, S. Guo, Z. Jiang, and Q. Lu, “Thermal-induced transverse-mode evolution in thermally guiding index-antiguided-core fiber,” Appl. Opt. 55(5), 1183–1189 (2016). [CrossRef]   [PubMed]  

13. L. Dong, “Approximate treatment of nonlinear waveguide equation in the regime of nonlinear self-focus,” J. Lightwave Technol. 26(20), 3476–3485 (2008). [CrossRef]  

14. L. Dong, “Formulation of a complex mode solver for arbitrary circular acoustic waveguides,” J. Lightwave Technol. 28, 3162–3175 (2010).

15. L. Dong, “Limits of stimulated Brillouin scattering suppression in optical fibers with transverse acoustic waveguide designs,” J. Lightwave Technol. 28, 3156–3161 (2010).

16. P. Yeh, A. Yariv, and E. Marom, “Theory of Bragg fiber,” J. Opt. Soc. Am. 68(9), 1196–1201 (1978). [CrossRef]  

17. D. C. Brown and H. J. Hoffman, “Thermal, stress, and thermal-optic effects in high average power double-clad silica fiber lasers,” IEEE J. Quantum Electron. 37(2), 207–217 (2001). [CrossRef]  

18. A. W. Snyder and J. D. Love, Optical Waveguide Theory (Chapman and Hall, 1983).

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 (8)

Fig. 1
Fig. 1 Mode collapse under the influence of thermal lensing. Equation (3) is evaluated for normalized frequency V = 3-8. A numerical simulation is also performed for comparison for fibers with core diameters of 10μm, 20μm and 30μm and NA of 0.06. The results are plotted as MFD for near-field MFD and eMFD for effective MFD, i.e. 2(Aeff/π) ½.
Fig. 2
Fig. 2 Comparison of the temperature profile from the numerical method used in this work and the parabolic solution for a uniform heat load. ΔT is set to be 0K at the cladding boundary. Core diameter is 20μm; cladding diameter is 400μm and NA is 0.06. The total heat load is 64w/m in both cases. The laser wavelength is at 1.03μm. The normalized heat load profile, i.e. normalized mode intensity profile, is also shown.
Fig. 3
Fig. 3 Typical run for the case in Fig. 2. Core diameter is 20μm and NA is 0.06. The total heat load is 64w/m. The laser wavelength is at 1.03μm. Relative changes are shown on the vertical axis on the right. The heat load is gradually increased over 30 steps.
Fig. 4
Fig. 4 Refractive index profiles under various heat loads. Core diameter is 20μm and NA is 0.06. The laser wavelength is at 1.03μm.
Fig. 5
Fig. 5 Simulated MFD for the fibers in [9]. Quantum defect heating is used to convert heat load to extracted power. Average heat load is used.
Fig. 6
Fig. 6 Simulated TMI nonlinear coupling coefficient for fibers with core diameters of 10μm, 15μm, 20μm, 25μm and 30μm respectively. The fiber NA is 0.06 and cladding diameter is 400μm. This study is conducted at a wavelength of 1.06μm. Contour lines indicate constant thermal load.
Fig. 7
Fig. 7 The LP01 and LP11 modes at 1.06μm in a fiber with a NA of 0.06 and a core diameter of 20μm without heat load and with a heat load of 1000w/m, i.e. ξQ0w02 = 2.19. The cladding diameter is 400μm.
Fig. 8
Fig. 8 Illustration of the scheme for solving the heat conduction equation with an arbitrary heat load profile with cylindrical symmetry.

Equations (36)

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

n(r)=n 0 (r) +k T ΔT(r)
ΔT(r)= Q 0 4πκ ρ 2 ( ρ 2 r 2 )
1 2 ξ Q 0 w 0 2 γ 6 +( U 2 +3ξ Q 0 ρ 2 + 9 2 ξ Q 0 w 0 2 ) γ 4 +3 ξ Q 0 ρ 2 γ 2 U 2 =0
ξ= k T k 0 2 n co 2 4πκ
t ψ(R)+{ w 2 k 0 2 n 0w 2 (R)[ 1+2 k T ΔT(R) ] w 2 β 2 m 2 R 2 }ψ(R)=0
t ψ 0 (R)+{ w 0 2 k 0 2 n 0w0 2 (R) w 0 2 β 0 2 m 2 R 2 } ψ 0 (R)=0
A [ w 0 2 ( k 0 2 n 0w0 2 β 0 2 ) ] w 2 ( k 0 2 n 0w 2 β 2 )ψ ψ 0 dA=2 k T w 2 k 0 2 A n 0w 2 ΔTdA+ A ψ 0 t ψψ t ψ 0 dA
w 2 β 2 w 0 2 β 0 2 =2 k T w 2 k 0 2 n T 2 k 0 2 ( w 0 2 n a 2 w 2 n b 2 )
w 2 = w 0 2 ( k 0 2 n a 2 β 0 2 )2 k T w 2 k 0 2 n T 2 k 0 2 n b 2 β 2
n a 2 = A n 0w0 2 (R)ψ(R) ψ 0 (R)dA A ψ (R)ψ 0 (R)dA   n b 2 = A n 0w 2 (R)ψ(R) ψ 0 (R)dA A ψ(R) ψ 0 (R)dA   n T 2 = A n 0w 2 (R)ΔT(R)ψ(R) ψ 0 (R)dA A ψ (R)ψ 0 (R)dA
t ψ(r)+{ k 0 2 n 0 2 (r)[ 1+2 k T ΔT(r) ] β 2 m 2 r 2 }ψ(R)=0
t ψ 0 (r)+{ k 0 2 n 0 2 (r) β 0 2 m 2 R 2 } ψ 0 (r)=0
β 2 β 0 2 =2 k T k 0 2 n Tr 2
n Tr 2 = A n 0 2 (r)ΔT(r)ψ(r) ψ 0 (r)dA A ψ(r) ψ 0 (r)dA
ψ(r)= e 1 2 r 2 w 2
n a 2 n co 2 n b 2 n co 2 n T 2 n co 2 Q 0 8πκ ρ 2 ( ρ 2 + w 2 2 )
n Tr 2 n co 2 Q 0 4πκ ρ 2 ( ρ 2 + 2 1 w 2 + 1 w 0 2 )
β 2 β 0 2 = k T k 0 2 n co 2 Q 0 2πκ ρ 2 ( ρ 2 + 2 1 w 2 + 1 w 0 2 )
1 2 ξ Q 0 w 0 2 γ 6 +( U 2 +3ξ Q 0 ρ 2 + 9 2 ξ Q 0 w 0 2 ) γ 4 +3 ξ Q 0 ρ 2 γ 2 U 2 =0
ξ= k T k 0 2 n co 2 4πκ
2 T(r) r 2 + 1 r T(r) r + Q(r) κ =0
T(r)= C 1 + C 2 ln(r)
T(r)=C Q 4κ r 2
T 1 (r)= T c Q 1 4κ r 1 2 r 1a 2 r 2
T 1a (r)= T c Q 1 4κ r 1 2 Q 1 2κ r 1 2 ln( r r 1a )
T 2 (r)= T 1a ( r 1 )+ Q 2 4κ r 2 2 r 1 2 r 2a 2 r 1 2 ( r 1 2 r 2 )
T 2a (r)= T 1a ( r 1 )+ Q 2 4κ ( r 1 2 r 2 2 ) Q 2 2κ r 2 2 r 1 2 r 2a 2 r 1 2 r 2a 2 ln( r r 2a )
r 2a 2 = r 1 2 + Q 2 Q 1 ( r 2 2 r 1 2 )
T 3 (r)= T 2a ( r 2 )+ Q 3 4κ r 3 2 r 2 2 r 3a 2 r 2 2 ( r 2 2 r 2 )
T 3a (r)= T 2a ( r 2 )+ Q 3 4κ ( r 2 2 r 3 2 ) Q 3 2κ r 3 2 r 2 2 r 3a 2 r 2 2 r 3a 2 ln( r r 3a )
r 3a 2 = r 2 2 + Q 3 Q 2 ( r 3 2 r 2 2 ) r 2a 2 r 1 2 r 2 2 r 1 2 r 2 2 r 2a 2
T N (r)= T (N1)a ( r N1 )+ Q N 4κ r N 2 r N1 2 r Na 2 r N1 2 ( r N1 2 r 2 )
T Na (r)= T (N1)a ( r N1 )+ Q N 4κ ( r N1 2 r N 2 ) Q N 2κ r N 2 r N1 2 r Na 2 r N1 2 r Na 2 ln( r r Na )
r Na 2 = r N1 2 + Q N Q N1 ( r N 2 r N1 2 ) r (N1)a 2 r N2 2 r N1 2 r N2 2 r N1 2 r (N1)a 2
T(r)= T (N1)a ( r N1 )+ Q N 4κ ( r N1 2 r N 2 ) Q N 2κ r N 2 r N1 2 r Na 2 r N1 2 r Na 2 ln( r r Na )
T 0 = T (N1)a ( r N1 )+ Q N 4κ ( r N1 2 r N 2 ) Q N 2κ r N 2 r N1 2 r Na 2 r N1 2 r Na 2 ln( b r Na )
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.