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

Mode instability in ytterbium-doped non-circular fibers

Open Access Open Access

Abstract

We present a theoretical study of transverse mode instability (TMI) in non-circular ytterbium-doped fibers including the rectangular core in a circular or D-shaped cladding. The D-shaped cladding is found efficient to suppress the TMI thanks to better heat dissipation, as compared to the circular cladding. However, the rectangular core does not suppress the TMI despite its better heat dissipation than a circular core counterpart. Although the temperature built in the rectangular core decreases with an increasing aspect ratio of the rectangular core, the low temperature does not benefit the TMI suppression. Instead, the TMI becomes stronger than its circular core counterpart. Our study reveals that the power coupling between two involved modes and gain saturation effect play a significant role in influencing the TMI. The power coupling strength is associated with the frequency offset between two modes, and it grows with an increasing aspect ratio of rectangular cores, suggesting the longer axis of rectangular core promotes the TMI.

© 2017 Optical Society of America

1. Introduction

High average power fiber laser with good beam quality is found as an attractive source in many applications for industrial, defense and scientific research [1]. Output powers of fiber lasers surpass bulk solid-state lasers thanks to its larger surface-to-active-volume ratio which efficiently extracts heat and makes robust single-mode operation [1,2]. J. W. Dawson et al. analyzed the power scalability of diffraction-limited fiber lasers and amplifiers [3], which shows that the highest achievable output power is determined by a combined effect of thermal lens and nonlinear scatterings such as stimulated Brillouin scattering (SBS) and stimulated Raman scattering (SRS). H. J. Otto et al. included transverse mode instability (TMI) in a similar analysis, and found that the TMI can occur before the combined limitation, hence setting another limitation for power scalability in commercial fibers with a V-parameter less than 14 [4]. The TMI was firstly reported in a large mode area (LMA) fiber amplifier in 2011 [5], and it is recognized as the major limitation for high average power scalability till now. Upon an onset of the TMI, the fundamental mode (FM) power transfers to a higher-order mode (HOM), and consequently the TMI degrades output beam quality as well as limits the FM power scalability. Soon after the observation of TMI phenomenon in high power fiber operation, C. Jauregui et al. suggested that the TMI is caused by the changed refractive index in core, which is induced by variations in the population of upper level lasing ions [6]. Shortly afterwards, A. Smith et al. considered the temporal evolution of the upper level ions population and temperature-induced index grating [7]. Now it is widely agreed that the TMI is caused by following three conditions [8]: a) FM and HOM interfere each other to produce an irradiance pattern that has a period equal to the beating length between modes; b) The irradiance pattern in fiber core induces different population inversion levels along the fiber, producing different heat generation through a quantum defect. The variation of generated heat leads to the formation of thermal grating via the thermo-optic effect. This temperature-induced refractive index grating accounts for the modal power transfer. c) A phase shift necessary for the TMI between irradiance pattern and thermal grating is induced by the thermal diffusion time across the core, frequency offset and different propagation constant between the associated modes.

A rectangular (or ribbon) core fiber design has been proposed for the effective FM area scaling above 1000 μm2 [9–14]. The asymmetric core offers bending insensitivity [11], robust higher-order mode area scaling [13], and better heat dissipation [9] than a circular core fiber, suggesting better TMI performance. However, there has been no investigation on TMI in this type of fiber, to the best of our knowledge. In this paper, we use a numerical model to present theoretical investigation of TMI in a rectangular core fiber. Compared to the semi-analytical model, the numerical model offers more flexible platform to adopt the asymmetric fiber geometry [15]. The influence of core shapes on the TMI is investigated by varying the core aspect ratio (AR.) and we present important parameters governing the TMI including the frequency offset between FM and HOM, TMI threshold pump power and temperature distribution in the non-circular fiber geometries. The benefit of better heat dissipation in the non-circular fibers is examined against the TMI threshold pump power, and we also employ a power coupling coefficient to attain more insight of the TMI behaviors in non-circular fibers. Hence, this paper contributes to the investigation of TMI in non-circular fibers, which hasn’t been examined in prior studies, with an emphasis on the influence of temperature, gain saturation effect, thermal lensing effect and power coupling between modes which is associated with frequency offset.

The paper is structured as follows: Section 2 describes the highly numeric model we employed. In section 3, we calculate the frequency offset that makes the most efficient power transfer from FM to HOM in rectangular core fibers as well as circular core fiber. In section 4, we compare the TMI threshold pump power between circular core and rectangular core fibers, and theoretically analyze the TMI behaviors in rectangular core fibers. The investigation on different cladding shapes is presented in section 5 and the summary of paper is provided in section 6.

2. TMI numerical model

The TMI in a fiber can be simulated via numerical models [15–18] or semi-analytical models [19–21]. A. Smith et al. proposed the highly numeric model in [15], and such model can handle a variety of fiber core shapes, refractive index profiles and doping profiles. The model intrinsically includes gain saturation effect and a thermal lensing effect that influence the TMI behavior. Furthermore, it allows including additional physical effects, such as photodarkening effect and bending effect. However, the highly numeric model needs longer computing time compared with other analytical models. K. R. Hansen et al. presented the semi-analytical model in [19,20] and the advantages of such model include shorter running time and ability to portray important physics of TMI with analytical expressions such as a power coupling coefficient. However, the semi-analytical model employs approximations and is usually formulated to solve a cylindrical symmetric fiber, or photonic crystal fiber [16]. To accommodate the asymmetric core designs of our interest and various contributing effects of the TMI, we use the highly numerical model, which is similar to the model in [15], with the fast Fourier transform (FFT) based beam propagation method (BPM). We briefly describe our model in the following, and here we only consider the LP01 and LP11 to define the TMI because the power transfer is strongest from LP01 to LP11 [8]. An input signal field is presented by:

Es(x,y,t)=P01E01(x,y)+P11E11(x,y)eiΔωt
where P01 is the LP01 mode input power and P11 is the LP11 mode input power, E01 and E11 are the normalized LP01 and LP11 mode amplitudes, respectively, Δω is the frequency offset between modes. The beam propagation equation is described as follows:
Es(x,y,z,t)z=i2kc2Es(x,y,z,t)i[k2(x,y,z,t)kc2]2kcEs(x,y,z,t)+g(x,y,z,t)Es(x,y,z,t)
where 2is the Laplacian operator in the x and y dimensions, g(x,y,z,t) is the laser gain, k(x,y,z,t) and kc are carrier wave numbers in the core and cladding, respectively. The wave number in the core and laser gain can be expressed as:
k(x,y,z,t)=ωcncore(x,y,z,t)c
g(x,y,z,t)=12[σsa+(σsa+σse)nu(x,y,z,t)]NYb(x,y)
where ωc is the carrier frequency, nclad is the refractive index of cladding, σsa and σse are the signal absorption and emission cross-sections, NYb(x,y) is the doping profile. The core refractive index during amplification, ncore(x,y,z,t), and upper state population, nu(x,y,z,t), are described through:
ncore(x,y,z,t)=ncore+dndTΔT(x,y,z,t)
nu(x,y,z,t)=Pp(z,t)σpahνpAp+Is(x,y,z,t)σsahνsPp(z,t)(σpa+σpe)hνpAp+Is(x,y,z,t)(σsa+σse)hνs+1τ
where ncore is a pristine refractive index of core, dn/dT is the thermo-optic coefficient, ΔT(x,y,z,t) is the increased temperature distribution in fiber, σpa and σpe are the pump absorption and emission cross-sections, Pp(z,t) is the pump power, Is(x,y,z,t) is the signal intensity, Ap is the pumping area, νs and νp are the signal and pump frequencies, h is the Planck constant and τ is the excited dopant ion lifetime.

We utilize the split-step FFT method to solve the BPM shown in Eq. (2), which has been proposed in [15,22–24]. In our simulation, each step size Δz consists of three sub-steps, and in the first and third sub-step, the field is advanced by Δz/2 in a homogeneous medium with a refractive index equal to the cladding index. In these two sub-steps, the BPM keeps only the diffractive term on right-hand-side as shown below and it is solved in the k space.

Es(x,y,z,t)z=i2kc2Es(x,y,z,t)
the field Es(x,y,z,t) is transformed to the k space by using the FFT presented as:
Es(x,y,z,t)=12πEs(kx,ky,z,t)eikxxeikyydkxdky
By substituting Eq. (8) into Eq. (7), we obtain:
Es(kx,ky,z,t)z=ikx22kcEs(kx,ky,z,t)+iky22kcEs(kx,ky,z,t)
Subsequently, the k space field Es(kx,ky,z,t) is advanced by Δz/2 consisting of a phase shift at each (kx,ky) plane as:
ϕ(kx,ky)=Δz2(kx2+ky2)2kc
Lastly, the IFFT is used to convert the k space field Es(kx,ky,z,t) back to Es(x,y,z,t).

In the second sub-step, the field is advanced by Δz, keeping only the phases induced by the thermally generated index plus guiding index, together with a signal gain on the right-hand-side of Eq. (2). Then the BPM for this step becomes:

Es(x,y,z,t)z=i[k2(x,y,z,t)kc2]2kcEs(x,y,z,t)+g(x,y,z,t)Es(x,y,z,t)
with the phase of the field advanced over Δz as:

ϕ(x,y,t)=Δzk2(x,y,t)kc22kc

To find the changed refractive index ncore(x,y,z,t) in Eq. (5), the increased temperature distribution ΔT(x,y,z,t) is calculated by using the Green’s function method presented in [15,25,26]. Additionally, we use a Dirichlet boundary condition for our thermal boundary, implying that temperature at the fiber surface is set at the room temperature, which is the same consideration as in [7,19]. When the temperature is fixed at the thermal boundary as indicated with dashed lines in Fig. 1, the Green’s function can be presented as:

G(x,y,ω|x',y')=n=0Fn(y,y')Pn(x,x',ω)
Fn(y,y')=1WKsin(nπWy)sin(nπWy')
Pn(x,x',ω)={exp[σn(2H|xx'|)]exp[σn(2Hxx')]+exp[σn|xx'|]exp[σn(x+x')]}/σn(1exp[2σnH])
σn2=(nπW)2+iωρCK
where 0 ≤ x ≤ H, 0 ≤ y ≤ W, and ω is the frequency of heat. The temperature can be calculated over the entire (x,y) grid with the contribution of periodically heated point (x’,y’) via the Green’s function. The heat equation to solve ΔT(x,y,z,t) is presented as follows:
ρCΔT(x,y,z,t)t=Q(x,y,z,t)+K(2ΔT(x,y,z,t)x2+2ΔT(x,y,z,t)y2)
where K is the thermal conductivity, ρ is the density, C is the heat capacity. The thermal load Q(x,y,z,t) in fiber core is expressed as:

 figure: Fig. 1

Fig. 1 Intensity of FM (LP01) and HOM (LP11) in (a) 50 μm circular core (Fiber 1), (b) square core 44.3 μm × 44.3 μm (AR. 1:1) (Fiber 2), (c) rectangular core 88.8 μm × 22.2 μm (AR. 4:1) (Fiber 3) and (d) rectangular core 140 μm × 14 μm (AR. 10:1) (Fiber 4). The dashed lines indicate the thermal boundary.

Download Full Size | PDF

Q(x,y,z,t)=NYb(x,y)[νpνsνp][σpa(σpa+σpe)nu(x,y,z,t)]Pp(z,t)Ap

The increased temperature distribution ΔT(x,y,z,t) can be solved using the Green’s function as:

ΔT(x,y,t)=Real[m=01x',y'q(x',y',ωm)G(x,y,ωm|x',y')exp(iωmt)]
We consider two frequency terms including ω = 0 and ω = Δω, which is sufficient for temperature calculation around the TMI threshold [15]. Higher frequency terms are only needed when calculation is required above the TMI threshold [15]. A temporal Fourier transform of thermal load Q(x’,y’,t) is presented as:

q(x',y',ωm)=ΔxΔyi=0Nt1Q(x',y',ti)exp(iωmti)

The time related pump power, Pp(z,t), FM power, P01(z,t) and HOM power, P11(z,t), are calculated at each fiber length z and time t using the following equations:

dPp(z,t)dz=Pp(z,t)Ap[(σpa+σpe)nu(x,y,z,t)σpa]NYb(x,y)dxdy
P01(z,t)=|Es(x,y,z,t)E01(x,y)dxdy[E01(x,y)]2dxdy|2P11(z,t)=|Es(x,y,z,t)E11(x,y)dxdy[E11(x,y)]2dxdy|2

Thus, the temperature distribution, pump power and individual modal power can be calculated along the fiber length. With Eqs. (1), (3) and (4), the TMI model can simulate an arbitrary shape of core, refractive index profile and doping profile. Additionally, the gain saturation effect is reflected by Eqs. (4) and (6), and it is influenced by the pump intensity and total signal intensity. The thermal lensing effect is also included in the first frequency term in Eq. (19).

3. Calculation of frequency offset for TMI

The phase shift between moving irradiance pattern and temperature grating is essential to provoke the power transfer from FM (LP01) to HOM (LP11). This phase shift is induced by the thermal diffusion time across the core and frequency offset between the modes. The most efficient power transfer in circular cores occurs at the phase shift where the frequency offset matches to the inversion of thermal diffusion time, thus the core radius [8], and such frequency offset agrees with the experimental result [27]. Now we present theoretical results on the most efficient frequency offset in rectangular cores which are asymmetric. We investigate the frequency offset in rectangular cores with different core AR. and this can be served as a good guideline to estimate significance of TMI in different core geometries. In the following, we compute the output power content of LP11 versus the frequency offset in rectangular cores with AR. of 1:1, 4:1 and 10:1 to find out the frequency offset that makes the most efficient power transfer from LP01 to LP11.

The parameters of circular core and rectangular core fibers used in this paper are listed in Table 1. Fiber 1 is a circular core fiber with the FM area of 1167 μm2, while Fiber 2 to Fiber 4 are rectangular core fibers with AR. of 1:1, 4:1 and 10:1, respectively, with the FM area slightly different from Fiber 1. The core area is kept nearly the same around 1960 μm2 for all fibers, which promises nearly the same total signal output power, pump power and pump efficiency (absorption). The overlap between modes is calculated by using E01(x,y)E11(x,y)dxdy.The pump and signal wavelengths are assumed at 976 and 1064 nm, respectively. We assume the Yb-doped aluminosilicate fiber, thus using signal emission and absorption cross-sections of 3.58 × 10−25 m2, and 6.00 × 10−27 m2, pump emission and absorption cross-sections of 1.87 × 10−24 m2, and 1.53 × 10−24 m2, respectively. Other fiber parameters used in the simulations are presented in the Table 1. For material properties, we use mass density of 2201 kg/m3, heat capacity of 702 J/(kg × K) and thermal conductivity of 1.38 W/(m × K). In the simulation, launched signal powers at LP01 and LP11 mode are set at 10 W, and 10−3 W, respectively, with a co-propagating pump beam (pumping radius fixed at 100 μm) in the cladding. We selected a bit higher LP11 mode input power than the quantum noise level. However, it does not affect our conclusions although it results in a bit low TMI threshold power. We note that even higher input power of 0.5 W was employed in [16]. Both the x and y grid spacings, Δx and Δy, are 1.56 μm. And a step size, Δz, is set at 10 μm. We select 64 time grid points for one time cycle (defined as 2π/Δω). Figure 1 shows the intensity of FM (LP01) and HOM (LP11) in circular core and rectangular core fibers. The dashed lines surrounding the circular cladding indicate the thermal boundary used to calculate the temperature distribution in fiber.

Tables Icon

Table 1. Fiber parameters used in simulation

We run simulation with the defined fiber parameters, following the aforementioned model, with varying the frequency offset between LP01 and LP11 modes. Figure 2 represents the calculated LP11 mode power content in the output power at various frequency offset values and we scale the peak power to a unity value for easy comparison. In the circular core fiber (Fiber 1), the LP11 power peaks at 1160 Hz, which is nearly the inversion of thermal diffusion time across the core (τ≈1.12r2, where r is the core radius in micrometers and τ is the thermal diffusion in microseconds [8]). In the rectangular cores, the peaked frequency offset shifts to lower values with the increasing AR. (or increasing longer axis and correspondingly decreasing shorter axis). Hence, the results suggest that the thermal diffusion time in the rectangular cores is determined by the longer axis rather than the shorter axis. In the rectangular cores, the modes are elongated along the longer axis [See Fig. 1], and their interaction is more intense along the longer axis. Therefore, the thermal diffusion time that accounts for the TMI follows the longer axis. Consequently, the core with AR. of 10 shows the lowest matched frequency offset at 260 Hz while the square core with AR. of 1 has the highest frequency offset of 1580 Hz due to its smallest core width among the investigated fibers [See Table 1]. The identified matched frequency offsets are used in the following simulations.

 figure: Fig. 2

Fig. 2 Normalized output LP11 mode power content versus frequency offset for 50 μm circular core fiber (Fiber 1), 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), 88.8 μm × 22.2 μm rectangular core (AR. 4:1) fiber (Fiber 3) and 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4).

Download Full Size | PDF

4. TMI in rectangular core fibers

The theoretical frequency offset for obtaining the maximum power transfer from LP01 to LP11 mode is in good agreement with the measured frequency in experiment [27]. Hence, we assume that the TMI becomes most efficient at the identified peak frequency offsets [see Fig. 2] even in the rectangular cores.

We firstly evaluate the efficiency of heat dissipation in rectangular cores. Figure 3 shows the increased temperature distribution (at t = 0) at the launch end when input pump power is assumed at 160 W within a 100 μm pumping radius. The signal powers are set to 10 W at LP01, and 10−3 W at LP11. The circular core fiber (Fiber 1) has its peak temperature of 1.58 K in the core, and the peak temperature reduces with larger AR. rectangular core, e.g. 1.486 K with AR. of 4, and 1.253 K with AR. of 10. The proximity to the cladding outer thermal boundary benefits the rectangular core with larger AR. for better heat dissipation. With the same reason, the square core 44.3 μm × 44.3 μm fiber (Fiber 2) is found to have the highest peak temperature of 1.644 K.

 figure: Fig. 3

Fig. 3 Increased temperature distribution (at t = 0) at the launch end in (a) 50 μm circular core fiber (Fiber 1), (b) 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), (c) 88.8 μm × 22.2 μm rectangular core (AR. 4:1) fiber (Fiber 3) and (d) 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4).

Download Full Size | PDF

Figure 4 represents the increased temperature profile (at t = 0) evolution along the fiber in a x-z plane (showing the longer axis of rectangular cores) with the same launched pump and signal powers as in Fig. 3. We calculate the temperature profile across the whole fiber structure, but only the core temperature is shown, as the increased temperature in fiber core can change the index profile as shown in Eq. (5), and then promotes the TMI. The temperature grows until the length at about 0.6 m, and gradually decreases along the fiber, and as expected, the larger AR. rectangular core builds lower temperature along the fiber. The Fiber 4 with AR. of 10 has 3.618 K of maximum temperature while the circular core fiber (Fiber 1) builds higher temperature of 4.103 K. We also note that the Fiber 2 (AR. of 1) builds the highest temperature of 4.241 K among the tested fibers, and the Fiber 3 with AR. of 4 reaches 3.957 K of maximum temperature. The results again manifest that rectangular core fiber with larger AR. has better thermal dissipation than circular core fiber. It is mainly because that larger AR. makes the fiber core (acts as heat source) edge closer to the cladding, therefore the heat can be more efficiently dissipated and the increased temperature in the core is suppressed. Hence, lower TMI is expected in the larger AR. rectangular core as compared to the circular core.

 figure: Fig. 4

Fig. 4 Increased temperature profile (at t = 0) evolution in the core along fiber length for (a) 50 μm circular core fiber (Fiber 1), (b) 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), (c) 88.8 μm × 22.2 μm rectangular core (AR. 4:1) fiber (Fiber 3) and (d) 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4). The same launched pump and signal powers are assumed as in Fig. 3.

Download Full Size | PDF

Figure 5 shows the TMI behaviors in the investigated fibers. Time averaged LP01 mode power, LP11 mode power, total signal power and pump power evolution are compared. All the tested fibers mark similar total signal output power around 150 W, however, the LP11 mode power grows differently. Interestingly, the larger AR. rectangular core suffers more significant TMI as evidenced with the higher output power of LP11 mode in Fig. 5, and this seems contradicting to the heat dissipation results. The output powers of pump and signal are listed in Table 2 for easy comparison. It is found that the LP11 mode constitutes 26.6% of the total signal power in Fiber 4 (AR. of 10) whereas the LP11 mode content is only 0.7% in the circular core (Fiber 1). Thus, the benefit of temperature reduction does not appear in the TMI. The pump efficiency indicates pump absorption, defined by (1-Ppout/Ppin) where Ppout and Ppin denote input and output pump powers, respectively. It is also noted from Table 2 that for the circular core (Fiber 1), the sum of LP01 and LP11 power is 149.64 W, which is consistent with total output power of 149.71 W. However, for the rectangular core (AR. 10:1) (Fiber 4), the sum of LP01 and LP11 power is only 147.05 W, less than total signal power of 148.6 W. We believe that this is mainly caused by a stronger thermal lensing effect in rectangular core (AR. 10:1) (Fiber 4). Due to the thermal lensing effect, other higher order modes could be populated [28] and some power of the LP01 mode is transferred to them. The thermal lensing effect on the TMI is presented in section 4.3.

 figure: Fig. 5

Fig. 5 Time averaged powers versus fiber length for (a) 50 μm circular core fiber (Fiber 1), (b) 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), (c) 88.8 μm × 22.2 μm rectangular core (AR. 4:1) fiber (Fiber 3) and (d) 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4). The same launched pump and signal powers are assumed as in Fig. 3.

Download Full Size | PDF

Tables Icon

Table 2. Output powers in the investigated fibers

We further investigate the TMI behaviors in the tested fibers while increasing the launched pump power, to evaluate the TMI threshold power. We consider the threshold pump power when output LP11 mode power grows to 10% of the total signal power. Figure 6 shows the output LP11 power content versus the input pump power. Again, the threshold pump power in Fiber 4 with AR. of 10 is much lower than in Fiber 1 despite the lower temperature in Fiber 4. The threshold power for Fiber 4 is only 61.7% (145 W) of Fiber 1 (238 W). In fact, the threshold pump power gets reduced with the increasing AR. despite better heat dissipation. We also note that Fiber 2 with AR. of 1 exhibits larger threshold by 14.9%, from 238 W to 273 W, than the circular core (Fiber 1), which contradicts to the temperature results again. Modes overlap or overlap between dopant and the modes might affect the TMI behavior. However, as shown in Table 1, the overlaps are nearly the same for all the fibers, and would have negligible influence on the TMI. Other parameters that can influence the TMI include signal effective mode areas which are substantially different in the tested fibers. In the following, we analyze the influence of different effective mode areas under the context of gain saturation effect on the TMI.

 figure: Fig. 6

Fig. 6 The output LP11 power content versus the input pump power for 50 μm circular core fiber (Fiber 1), 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), 88.8 μm × 22.2 μm rectangular core (AR. 4:1) fiber (Fiber 3) and 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4). The same launched signal powers are assumed as in Fig. 3.

Download Full Size | PDF

4.1 Gain saturation effect

In the Table 1, the Fiber 4 has the largest signal effective mode area of 1368 μm2, which is 17.2% larger than Fiber 1 having 1167 μm2 of the signal mode area. We also note that Fiber 2 has the smallest signal mode area of 1083 μm2 and the signal mode area of Fiber 3 is positioned between the Fiber 1 and Fiber 4. A higher intensity of signal or lower intensity of pump expedites depletion of excited Yb ions, and brings in the gain saturation effect more significantly. The stronger gain saturation is found efficient to suppress the TMI, hence a fiber with smaller signal effective mode area will provide better TMI suppression [29]. In our simulation, Fiber 4 contains the largest signal mode area, thus being expected for lower signal intensity and weaker gain saturation effect. We test the gain saturation effect as shown in Fig. 7. The upper state population (n2) profiles and the normalized total signal intensity (scaled to unity at peak) at the input end are presented for Fiber 1, Fiber 2 and Fiber 4 when pump power is assumed at 150 W. Here we consider the overlap between n2 and signal intensity profiles to evaluate the gain saturation in that the smaller overlap means the stronger gain saturation effect and vice versa. The overlap calculated for Fiber 1, Fiber 2 and Fiber 4 are 0.9962, 0.9937 and 0.9987, respectively, when a pumping radius is 100 μm the same as in previous plots. Therefore, the overlap is calculated at the maximum in Fiber 4 and the minimum in Fiber 2, and we also note that the Fiber 1 has the overlap in-between the other two fibers. Therefore, it is expected that the significance of the TMI follows the significance of the overlap, i.e. lower TMI threshold power for Fiber 4 and higher threshold power for Fiber 2. This seems consistent with our simulation results shown in Fig. 6.

 figure: Fig. 7

Fig. 7 The normalized signal intensity and upper state population (n2) profiles at the input end for (a) 50 μm circular core fiber (Fiber 1), (b) 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), (c) 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4). Insets enlarge the selected portion to emphasize the overlap differences by adjusting pump intensities. The same launched pump and signal powers are assumed as in Fig. 3. PR represents the pumping radius.

Download Full Size | PDF

It is then interesting to examine the TMI when the gain saturation effect becomes equivalent in all fibers. To achieve the same level of saturation effect in all tested fibers, we adjust the pump intensity by changing the pumping radius in Fiber 2 and Fiber 4 as illustrated in Figs. 7(b) and 7(c). The pump intensity adjustment leads to the changes in overlap to 0.9965 and 0.9961 in Fiber 2 and Fiber 4, respectively, and the overlap in Fiber 1 is maintained as 0.9962. Then, all the fibers offer nearly the same overlaps, thus the same level of saturation effect.

Consequently, the TMI behavior at the same level of saturation effect is presented in Fig. 8. As expected, the TMI threshold power is reduced in Fiber 2 when difference in the saturation effect is lifted [See Fig. 6 for comparison]. Fiber 2 follows the same TMI behavior as the circular core fiber (Fiber 1), albeit not completely the same. Its TMI threshold pump power drops from 273 W to 245 W, which is quite close to the threshold power of Fiber 1 at 238 W. Therefore, we conclude that the stronger gain saturation in Fiber 2 plays a key role in suppressing the TMI as compared to the Fiber 1. In contrast, Fiber 4 does not follow the expectation even when the difference of gain saturation effect is removed. Although the threshold pump power is improved from 145 W to 182 W with the enhanced saturation effect, the fiber still exhibits the strongest TMI despite its lowest temperature even at the same level of saturation.

 figure: Fig. 8

Fig. 8 The output LP11 power content versus the input pump power for 50 μm circular core fiber (Fiber 1), 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), and 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4). The same launched pump and signal powers are assumed as in Fig. 3.

Download Full Size | PDF

4.2 Power coupling strength

The TMI is viewed as the stimulated thermal Rayleigh scattering (STRS) that induces a power coupling between supported modes in an optical fiber, and the power coupling strength can be evaluated by introducing a power coupling coefficient. It is reported that the coupling coefficient of STRS is determined by a few factors including the doping area in a core, temperature built by a quantum defect, an overlap of the supported modes and the frequency offset between two modes [8,19]. For instance, the power coupling coefficient can be described as [19]:

χ1,2(Δω)=dndTk2Kβ1,2Im[A(Δω)](1λsλp)
where k is the vacuum wavenumber, β1 and β2 are the propagation constant for LP01 mode and LP11 mode. Therefore, the coupling coefficient is mainly determined by the quantity Im(A) because other terms are related to material constants and waveguide parameters. The Im(A) is described as
Im(A(Δω))=ΩE01(x,y)E11(x,y)×(ΩdIm(G(x,y,x',y',Δω))E01(x',y')E11(x',y')dx'dy')dxdy
where Ω and Ωd denote the whole fiber cross section and doped cross section, G is the Green’s function used to solve the temperature distribution shown in Eq. (17), and the expression of G is presented in Eqs. (13)-(16).

The power coupling coefficient (χ1 nearly equals to χ2) is calculated as 0.4213 W−1 for Fiber 4, which is much larger than 0.1167 W−1 for Fiber 1. The larger χ implies a stronger power coupling between modes, which manifests the stronger TMI. Thus, we conclude that Fiber 4 suffers more significant TMI as shown in Fig. 6 due to its 3.6 times stronger coupling strength than Fiber 1. The benefit of the lower temperature in Fiber 4 is overshadowed by the strong coupling coefficient, and does not help to suppress the TMI. Interestingly, the coupling coefficient is a function of the frequency offset, and we have shown that the frequency offset is mainly determined by the longer axis in a rectangular core fiber. Therefore, the coupling strength in a rectangular core fiber is similar to the circular core with a core diameter same as the longer axis of the rectangular core if the differences in gain saturation effect and increased temperature distribution are insignificant. Additionally, the quantity power coupling coefficient in Fiber 2 is 0.1042 W−1, which is a little smaller than 0.1167 W−1 in Fiber 1, hence implying that the TMI is a little weaker in Fiber 2 compared with that in Fiber 1 as shown in Fig. 8.

4.3 Thermal lensing effect

When the signal power is mostly confined to the fundamental mode, the temperature develops its profile following the fundamental mode, resulting in temperature peak at the center of fiber core. This in turn causes the thermal lensing effect in a fiber [27]. Figure 9 presents the signal effective area as a function of propagation length, z, when the same launched pump and signal powers are assumed as in Fig. 3 and pumping radius fixed at 100 μm. It is clear to see that in both fibers, the signal effective area decreases until about z = 0.5 m, and subsequently increases to recover the original effective area at the output end. This is attributed to the thermal lensing effect, which corresponds to the temperature profile evolution as shown in Fig. 4. The static temperature becomes its maximum at ~0.5 m, hence most significant thermal lensing effect. The maximum signal effective area reduction in a circular core (Fiber 1) amounts to 2.23%, which is much less than 9.58% in Fiber 4 with AR. of 10, implying that rectangular core with larger AR (or wider core) generates stronger thermal lensing. As the thermal lensing effect mainly pushes the fundamental mode to the core center, the modal overlap gets lowered, which leads to an increase in TMI threshold power [30]. Additionally, the reduced signal effective mode area caused by thermal lensing can enhance the gain saturation effect, thus leading to TMI suppression again. Therefore, the stronger thermal lensing effect in a rectangular core fiber with larger AR. helps increase the TMI threshold power as compared to a circular core fiber. Even so, the stronger power coupling between modes in rectangular core fiber with larger AR. overshadows the thermal lensing effect, to make the TMI stronger in the rectangular cores. We also note that Fiber 4 exhibits more significant TMI along the entire fiber than Fiber 1 even at the point of the maximum thermal lensing effect as presented in Fig. 5.

 figure: Fig. 9

Fig. 9 The signal effective area as a function of propagation length, z, for (a) 50 μm circular core fiber (Fiber 1) and (b) 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) when the same launched pump and signal powers are assumed as in Fig. 3.

Download Full Size | PDF

4.4 TMI threshold power versus core area and effective mode area

The relationship between threshold pump power and core area is presented in Fig. 10. We assume the core areas of 1960 μm2, 2940 μm2, 3920 μm2, 5880 μm2 and 7840 μm2, for circular, square core (AR. 1:1), rectangular core (AR. 4:1) and rectangular core (AR. 10:1) fibers. As expected from the previous results, it is clear to see that when the core area is same, the square core (AR. 1:1) fiber has the highest threshold pump power due to the strongest gain saturation effect and the weakest power coupling strength. On the other hand, the rectangular core (AR. 10:1) fiber suffers the smallest threshold power, which is mainly caused by the strongest power coupling and then the weakest gain saturation effect among the tested fibers. It is also noted that in all the tested fibers, the threshold power decreases with the increasing core area by weakening the gain saturation effect.

 figure: Fig. 10

Fig. 10 The threshold pump power versus core area for circular core, square core (AR. 1:1), rectangular core (AR. 4:1) and rectangular core (AR. 10:1) fibers. The same launched signal powers are assumed as in Fig. 3.

Download Full Size | PDF

Figure 11 shows the threshold pump power (left axis) and power coupling coefficient χ (right axis) for circular core, square core (AR. 1:1), rectangular cores with AR. of 4:1 and AR. of 10:1 when the input signal effective mode area is fixed at 1170 μm2, 1750 μm2, 2300 μm2, 3500 μm2 and 4650 μm2. As seen in Fig. 11, when the effective mode area is fixed, the threshold power in a circular core fiber is very close to its square core counterpart, and the slightly higher TMI threshold power in the square fiber is attributed to the slightly smaller coupling strength, χ. That is to say, the stronger gain saturation effect in square core (AR. 1:1) fiber is the main factor contributing to the higher threshold power than the circular core fiber shown in Fig. 10. Therefore, we confirm that with the same gain saturation effect, the TMI threshold power can be mainly decided by the power coupling strength. Similarly, the TMI threshold power is decreasing with the increasing AR. in rectangular cores. Even with the same effective mode area (thus nearly same saturation effect), the larger AR. core sees stronger TMI due to the stronger coupling strength. Apparently, the stronger coupling is caused by the lower frequency offset in rectangular cores with larger AR. Furthermore, the coupling strength gets stronger with the increasing effective mode area in all the tested fibers because of the increasing core size.

 figure: Fig. 11

Fig. 11 The threshold pump power (left axis) and power coupling coefficient χ (right axis) versus effective mode area for circular core, rectangular core (AR. 1:1), rectangular core (AR. 4:1) and rectangular core (AR. 10:1) fibers. The same launched signal powers are assumed as in Fig. 3.

Download Full Size | PDF

5. TMI in D-shaped cladding fibers

We now move on to investigate the TMI in a D-shaped cladding fiber when the core is in a circular or rectangular shape. The purpose of this work is to confirm the influence of temperature alone on the TMI when differences in the gain saturation effect and the power coupling strength become negligible. We compare the TMI between D-shaped cladding and circular cladding fibers, and select the core designs of Fiber 1 and Fiber 4 for the investigation. Figure 12 represents the investigated fiber structures having a D-shaped cladding with 400 μm and 200 μm in long and short axes, respectively. The dashed lines represent the thermal boundary position in D-shaped cladding fibers, and the Dirichlet boundary condition is assumed as in the circular cladding fibers. In both the circular and D-shaped cladding fibers, we assume the same pumping radius of 100 μm to keep the gain saturation effect the same. The input signal powers at LP01 and LP11 mode are 10 W and 10−3 W, respectively. The parameters of fiber core in D-shaped cladding are kept the same as listed in Table 1. As the frequency offset is mainly determined by the thermal diffusion time (heat transfer from core center to core edge) [8] or signal effective mode area [27], hence the circular core and rectangular core in D-shaped cladding fibers have the nearly same frequency offset as in circular cladding fibers at 1160 Hz and 260 Hz, respectively.

 figure: Fig. 12

Fig. 12 Intensity of (a) LP01 and (b) LP11 mode in the 50 μm circular core fiber (Fiber 1) with D-shaped cladding, (c) LP01 and (d) LP11 in the 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) with D-shaped cladding. The same core parameters are used as listed in Table 1. The dashed lines indicate the thermal boundary position.

Download Full Size | PDF

5.1 TMI in circular core D-shaped cladding fiber

The increased temperature distributions (at t = 0) at the launch end of Fiber 1 with the circular and D-shaped cladding are shown in Fig. 13. Input pump power of 250W is assumed while the modal input powers are kept the same as in Fig. 3. Clearly, the D-shaped cladding helps to reduce the temperature from 1.681 K in a circular cladding to 1.384 K in the D-shaped cladding. The temperature suppression is achieved by the cladding geometry, i.e. the circular core (acts as a heat source) gets closer to the cladding boundary (acts as a heat sink) in the D-shaped cladding. The corresponding increased temperature profile evolution (at t = 0) along the fiber is presented in Fig. 14. The maximum increased temperature is reduced by about 14.6% from 6.217 K in the circular cladding to 5.31 K in the D-shaped cladding, which reconfirms that the D-shaped cladding fiber provides more efficient heat dissipation. Hence, better TMI suppression is expected in the D-shaped cladding.

 figure: Fig. 13

Fig. 13 Increased temperature distribution (at t = 0) within the whole core and cladding region at the launch end in 50 μm circular core (Fiber 1) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers are assumed as in Fig. 3. The input pump power is 250 W with 100 μm pumping radius.

Download Full Size | PDF

 figure: Fig. 14

Fig. 14 Increased temperature profile evolution (at t = 0) within the core along fiber in 50 μm circular core (Fiber 1) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers, pump power and pumping radius are assumed as in Fig. 13.

Download Full Size | PDF

Our numerical simulation results of time averaged LP01 and LP11 mode power, total signal power and pump power evolution are shown in Fig. 15. The higher output power of LP11 mode in the circular cladding fiber is due to the higher temperature compared with the D-shaped cladding fiber. Hence, the results confirm that better heat dissipation helps suppress the TMI when the same gain saturation effect and power coupling strength are applied. The TMI threshold pump power together with the maximum increased core temperature of both cladding shapes are examined by increasing the pump power, and the results are shown in Fig. 16. The threshold pump power in the D-shaped cladding is reduced by 17% from 275 W to 235 W as compared to the circular cladding. This improvement is attributed to the temperature decrease by about 15% as indicated in the right axis in Fig. 16.

 figure: Fig. 15

Fig. 15 Time averaged powers versus fiber length for 50 μm circular core (Fiber 1) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers, pump power and pumping radius are assumed as in Fig. 13.

Download Full Size | PDF

 figure: Fig. 16

Fig. 16 The output LP11 power content (left axis) and maximum increased temperature in the core (right axis) versus the input pump power for 50 μm circular core (Fiber 1) with circular cladding and D-shaped cladding. The same launched signal powers and pumping radius are assumed as in Fig. 13.

Download Full Size | PDF

5.2 TMI in rectangular core D-shaped cladding fiber

We extend our investigation to the rectangular core with AR. of 10 (Fiber 4) in a circular cladding and a D-shaped cladding. Figure 17 shows the increased temperature distribution (at t = 0) at the launch end for both fibers when the input pump power is 140 W with 100 μm pumping radius. As before, the D-shaped cladding facilitates the heat dissipation, resulting in lower temperature of 0.96 K than 1.23 K in the circular cladding. The increased temperature profile evolution (at t = 0) within the fiber core is also presented for both fibers in Fig. 18. The simulation result confirms that D-shaped cladding reaches lower temperature of 2.05 K than 2.72 K in the circular cladding, which is about 24.7% reduction. The larger reduction than circular core (Fiber 1) suggests that the rectangular core (Fiber 4) offers more efficient heat dissipation than circular core in the D-shaped cladding.

 figure: Fig. 17

Fig. 17 Increased temperature distribution (at t = 0) at the launch end within the whole core and cladding region in 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers are assumed as in Fig. 3. The input pump power is 140 W with 100 μm pumping radius.

Download Full Size | PDF

 figure: Fig. 18

Fig. 18 Increased temperature profile evolution (at t = 0) in the core along fiber for 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers and pumping radius are assumed as in Fig. 17.

Download Full Size | PDF

Figure 19 presents the time averaged LP01 and LP11 mode power, total signal power and pump power evolution in both cladding shapes. The output LP11 mode power in the D-shaped cladding fiber grows to 1.494 W shown in Fig. 19(b), which is much less than 10 W in the circular cladding fiber presented in Fig. 19(a). Thus, the efficient heat dissipation is found helpful to suppress the TMI when other factors such as the saturation effect and coupling strength are same in both fibers.

 figure: Fig. 19

Fig. 19 Time averaged powers versus fiber length for 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers and pumping radius are assumed as in Fig. 17.

Download Full Size | PDF

Figure 20 shows the output LP11 power content (left axis) and maximum increased temperature in the core (right axis) versus input pump power. Apparently, the maximum increased temperature in the D-shaped cladding fiber drops by about 24.7% compared to the circular cladding fiber. Consequently, the threshold pump power in the D-shaped cladding fiber increases by 29% compared to the circular cladding fiber, from 145 W to 187 W. We note that the threshold power of the rectangular core (AR. 10:1) D-shaped cladding fiber is much lower than the circular core circular cladding fiber (Fiber 1) (187 W versus 235 W) due to the stronger power coupling and weaker gain saturation effect in the rectangular core. We also note that the temperature reduction by introducing the D-shaped cladding in the rectangular core amounts to 24.7%, which is larger than the temperature drop of 14.6% in the circular core counterparts. Despite the larger temperature reduction in the rectangular core induced by the D-shaped cladding, the TMI does not follow the temperature reduction, but is mainly determined by the gain saturation effect and power coupling strength.

 figure: Fig. 20

Fig. 20 The output LP11 power content (left axis) and maximum increased temperature in the core (right axis) versus the input pump power for 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) with circular cladding and D-shaped cladding. The same launched signal powers and pumping radius are assumed as in Fig. 17.

Download Full Size | PDF

6. Conclusion

In conclusion, we present theoretical investigation of the TMI in non-circular fibers including rectangular core fibers and D-shaped cladding fibers for the first time, to the best of our knowledge. Our results reveal that the temperature built in a fiber is not solely responsible for the TMI, but the gain saturation effect, coupling strength and thermal lensing effect also play a significant role in governing the TMI behavior. Furthermore, the coupling strength is a function of the frequency offset, which is mainly determined by the longer axis in a rectangular core fiber. Thus, the power coupling strength in a rectangular core fiber can be nearly as strong as a circular core fiber with the same core diameter as the longer axis of the rectangular core. On the other hand, when the same core area (or signal effective area) is applied to both the circular and the rectangular cores, the longer axis in the rectangular core is larger than the core diameter in the circular core. Consequently, the TMI threshold power becomes higher in the circular core because of the weaker power coupling strength due to higher frequency offset. Temperature reduction by shaping the core does not always suppress the TMI, although shaping the cladding can suppress the TMI via better heat dissipation.

In our investigation, at the fixed core area, the rectangular core (AR. 10:1) fiber offers the most efficient heat dissipation (24.6% better than a circular core) among the circular core, square core (AR. 1:1), and rectangular cores with AR. of 4 and 10. However, rectangular core (AR. 10:1) fiber suffers the worst TMI among the tested fibers, which drops the threshold pump power by 38.3%, from 238 W to 145 W, in comparison with circular core fiber. We find this is mainly caused by the stonger power coupling strength and then the weaker gain saturation effect than other types of fibers. Additionally, we also simulate the TMI in D-shaped cladding fibers, and find that such cladding structure offers more efficient heat dissipation, which suppresses the TMI and increases the threshold pump power when gain saturation effect and power coupling strength are the same. For instance, the D-shaped cladding circular core and retangular core (AR. 10:1) fibers reduce the temperature by 15% and 25%, respectively, and increase the threshold pump power by 17% and 35.2% when compared to circular cladding counterparts. Despite this better heat dissipation, the rectangular core (AR. 10:1) D-shaped cladding fiber suffers much more significant TMI with 189 W threshold power than its circular core circular cladding counterpart having 235 W threshold power, which is opposite to the temperature increase. Our investigation reconfirms that the significance of TMI is not simply predicted by temperature, as the influence of temperature on TMI can be overshadowed by gain saturation effect and power coupling strength between modes.

Funding

A*STAR; KEIT.

Acknowledgment

This work is supported by A*STAR through Advanced Optics Engineering programme. S. Yoo acknowledges support of KEIT through global research programme.

References and links

1. D. J. Richardson, J. Nilsson, and W. A. Clarkson, “High power fiber lasers: current status and future perspectives,” J. Opt. Soc. Am. B 27(11), B63–B92 (2010). [CrossRef]  

2. C. Jauregui, J. Limpert, and A. Tünnermann, “High-power fiber lasers,” Nat. Photonics 7(11), 861–867 (2013). [CrossRef]  

3. 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]  

4. H. J. Otto, C. Jauregui, J. Limpert, and A. Tunnermann, “Average power limit of fiber-laser systems with nearly diffraction-limited beam quality,” Proc. SPIE 9728, 97280E (2016).

5. 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]  

6. C. Jauregui, T. Eidam, J. Limpert, and A. Tünnermann, “The impact of modal interference on the beam quality of high-power fiber amplifiers,” Opt. Express 19(4), 3258–3271 (2011). [CrossRef]   [PubMed]  

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

8. A. V. Smith and J. J. Smith, “Overview of a steady-periodic model of modal instability in fiber amplifiers,” IEEE J. Sel. Top. Quantum Electron. 20(5), 472–483 (2014). [CrossRef]  

9. D. A. Rockwell, V. V. Shkunov, and J. R. Marciante, “Semi-guiding high-aspect-ratio core (SHARC) fiber providing single-mode operation and an ultra-large core area in a compact coilable package,” Opt. Express 19(15), 14746–14762 (2011). [CrossRef]   [PubMed]  

10. D. R. Drachenberg, M. J. Messerly, P. H. Pax, A. K. Sridharan, J. B. Tassano, and J. W. Dawson, “Yb3+ doped ribbon fiber for high-average power lasers and amplifiers,” Proc. SPIE 8961, 89610T (2014). [CrossRef]  

11. S. Yoo, J. Ji, X. Wu, S. Raghurman, D. Ho, N. Xia, J. Sahu, and J. Nilsson, “Mode area scalability in rectangular core fiber,” in Proceedings of IEEE Conference on Photonics (IEEE, 2015), pp. 323–324. [CrossRef]  

12. T. S. Saini, A. Kumar, Y. Kalra, and R. K. Sinha, “Rectangular-core large-mode-area photonic crystal fiber for high power applications: design and analysis,” Appl. Opt. 55(15), 4095–4100 (2016). [CrossRef]  

13. D. Drachenberg, M. Messerly, P. Pax, A. Sridharan, J. Tassano, and J. Dawson, “First multi-watt ribbon fiber oscillator in a high order mode,” Opt. Express 21(15), 18089–18096 (2013). [CrossRef]   [PubMed]  

14. A. K. Sridharan, P. H. Pax, J. E. Heebner, D. R. Drachenberg, J. P. Armstrong, and J. W. Dawson, “Mode-converters for rectangular-core fiber amplifiers to achieve diffraction-limited power scaling,” Opt. Express 20(27), 28792–28800 (2012). [CrossRef]   [PubMed]  

15. A. V. Smith and J. J. Smith, “Steady-periodic method for modeling mode instability in fiber amplifiers,” Opt. Express 21(3), 2606–2623 (2013). [CrossRef]   [PubMed]  

16. B. G. Ward, “Modeling of transient modal instability in fiber amplifiers,” Opt. Express 21(10), 12053–12067 (2013). [CrossRef]   [PubMed]  

17. B. Ward, C. Robin, and I. Dajani, “Origin of thermal modal instabilities in large mode area fiber amplifiers,” Opt. Express 20(10), 11407–11422 (2012). [CrossRef]   [PubMed]  

18. S. Naderi, I. Dajani, T. Madden, and C. Robin, “Investigations of modal instabilities in fiber amplifiers through detailed numerical simulations,” Opt. Express 21(13), 16111–16129 (2013). [CrossRef]   [PubMed]  

19. K. R. Hansen, T. T. Alkeskjold, J. Broeng, and J. Lægsgaard, “Thermally induced mode coupling in rare-earth doped fiber amplifiers,” Opt. Lett. 37(12), 2382–2384 (2012). [CrossRef]   [PubMed]  

20. K. R. Hansen, T. T. Alkeskjold, J. Broeng, and J. Lægsgaard, “Theoretical analysis of mode instability in high-power fiber amplifiers,” Opt. Express 21(2), 1944–1971 (2013). [CrossRef]   [PubMed]  

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

22. M. D. Feit and J. A. Fleck Jr., “Computation of mode eigenfunctions in graded-index optical fibers by the propagating beam method,” Appl. Opt. 19(13), 2240–2246 (1980). [CrossRef]   [PubMed]  

23. M. D. Feit and J. A. Fleck Jr., “Computation of mode properties in optical fiber waveguides by a propagating beam method,” Appl. Opt. 19(7), 1154–1164 (1980). [CrossRef]   [PubMed]  

24. M. D. Feit and J. A. Fleck Jr., “Light propagation in graded-index optical fibers,” Appl. Opt. 17(24), 3990–3998 (1978). [CrossRef]   [PubMed]  

25. K. D. Cole and D. H. Yen, “Green’s functions, temperature and heat flux in the rectangle,” J. Heat Transfer 44(20), 3883–3894 (2001). [CrossRef]  

26. K. D. Cole, “Steady-periodic Green’s functions and thermal-measurement applications in rectangular coordinates,” J. Heat Transfer 128(7), 709–716 (2006). [CrossRef]  

27. A. V. Smith and J. J. Smith, “Frequency dependence of mode coupling gain in Yb doped fiber amplifiers due to stimulated thermal Rayleigh scattering,” arXiv: 1301.4277 [physics. optics] (2013).

28. W. W. Ke, X. J. Wang, X. F. Bao, and X. J. Shu, “Thermally induced mode distortion and its limit to power scaling of fiber lasers,” Opt. Express 21(12), 14272–14281 (2013). [CrossRef]   [PubMed]  

29. A. V. Smith and J. J. Smith, “Increasing mode instability thresholds of fiber amplifiers by gain saturation,” Opt. Express 21(13), 15168–15182 (2013). [CrossRef]   [PubMed]  

30. L. Dong, “Thermal lensing in optical fibers,” Opt. Express 24(17), 19841–19852 (2016). [CrossRef]   [PubMed]  

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 (20)

Fig. 1
Fig. 1 Intensity of FM (LP01) and HOM (LP11) in (a) 50 μm circular core (Fiber 1), (b) square core 44.3 μm × 44.3 μm (AR. 1:1) (Fiber 2), (c) rectangular core 88.8 μm × 22.2 μm (AR. 4:1) (Fiber 3) and (d) rectangular core 140 μm × 14 μm (AR. 10:1) (Fiber 4). The dashed lines indicate the thermal boundary.
Fig. 2
Fig. 2 Normalized output LP11 mode power content versus frequency offset for 50 μm circular core fiber (Fiber 1), 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), 88.8 μm × 22.2 μm rectangular core (AR. 4:1) fiber (Fiber 3) and 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4).
Fig. 3
Fig. 3 Increased temperature distribution (at t = 0) at the launch end in (a) 50 μm circular core fiber (Fiber 1), (b) 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), (c) 88.8 μm × 22.2 μm rectangular core (AR. 4:1) fiber (Fiber 3) and (d) 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4).
Fig. 4
Fig. 4 Increased temperature profile (at t = 0) evolution in the core along fiber length for (a) 50 μm circular core fiber (Fiber 1), (b) 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), (c) 88.8 μm × 22.2 μm rectangular core (AR. 4:1) fiber (Fiber 3) and (d) 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4). The same launched pump and signal powers are assumed as in Fig. 3.
Fig. 5
Fig. 5 Time averaged powers versus fiber length for (a) 50 μm circular core fiber (Fiber 1), (b) 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), (c) 88.8 μm × 22.2 μm rectangular core (AR. 4:1) fiber (Fiber 3) and (d) 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4). The same launched pump and signal powers are assumed as in Fig. 3.
Fig. 6
Fig. 6 The output LP11 power content versus the input pump power for 50 μm circular core fiber (Fiber 1), 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), 88.8 μm × 22.2 μm rectangular core (AR. 4:1) fiber (Fiber 3) and 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4). The same launched signal powers are assumed as in Fig. 3.
Fig. 7
Fig. 7 The normalized signal intensity and upper state population (n2) profiles at the input end for (a) 50 μm circular core fiber (Fiber 1), (b) 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), (c) 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4). Insets enlarge the selected portion to emphasize the overlap differences by adjusting pump intensities. The same launched pump and signal powers are assumed as in Fig. 3. PR represents the pumping radius.
Fig. 8
Fig. 8 The output LP11 power content versus the input pump power for 50 μm circular core fiber (Fiber 1), 44.3 μm × 44.3 μm square core (AR. 1:1) fiber (Fiber 2), and 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4). The same launched pump and signal powers are assumed as in Fig. 3.
Fig. 9
Fig. 9 The signal effective area as a function of propagation length, z, for (a) 50 μm circular core fiber (Fiber 1) and (b) 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) when the same launched pump and signal powers are assumed as in Fig. 3.
Fig. 10
Fig. 10 The threshold pump power versus core area for circular core, square core (AR. 1:1), rectangular core (AR. 4:1) and rectangular core (AR. 10:1) fibers. The same launched signal powers are assumed as in Fig. 3.
Fig. 11
Fig. 11 The threshold pump power (left axis) and power coupling coefficient χ (right axis) versus effective mode area for circular core, rectangular core (AR. 1:1), rectangular core (AR. 4:1) and rectangular core (AR. 10:1) fibers. The same launched signal powers are assumed as in Fig. 3.
Fig. 12
Fig. 12 Intensity of (a) LP01 and (b) LP11 mode in the 50 μm circular core fiber (Fiber 1) with D-shaped cladding, (c) LP01 and (d) LP11 in the 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) with D-shaped cladding. The same core parameters are used as listed in Table 1. The dashed lines indicate the thermal boundary position.
Fig. 13
Fig. 13 Increased temperature distribution (at t = 0) within the whole core and cladding region at the launch end in 50 μm circular core (Fiber 1) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers are assumed as in Fig. 3. The input pump power is 250 W with 100 μm pumping radius.
Fig. 14
Fig. 14 Increased temperature profile evolution (at t = 0) within the core along fiber in 50 μm circular core (Fiber 1) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers, pump power and pumping radius are assumed as in Fig. 13.
Fig. 15
Fig. 15 Time averaged powers versus fiber length for 50 μm circular core (Fiber 1) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers, pump power and pumping radius are assumed as in Fig. 13.
Fig. 16
Fig. 16 The output LP11 power content (left axis) and maximum increased temperature in the core (right axis) versus the input pump power for 50 μm circular core (Fiber 1) with circular cladding and D-shaped cladding. The same launched signal powers and pumping radius are assumed as in Fig. 13.
Fig. 17
Fig. 17 Increased temperature distribution (at t = 0) at the launch end within the whole core and cladding region in 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers are assumed as in Fig. 3. The input pump power is 140 W with 100 μm pumping radius.
Fig. 18
Fig. 18 Increased temperature profile evolution (at t = 0) in the core along fiber for 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers and pumping radius are assumed as in Fig. 17.
Fig. 19
Fig. 19 Time averaged powers versus fiber length for 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) with (a) circular cladding and (b) D-shaped cladding. The same launched signal powers and pumping radius are assumed as in Fig. 17.
Fig. 20
Fig. 20 The output LP11 power content (left axis) and maximum increased temperature in the core (right axis) versus the input pump power for 140 μm × 14 μm rectangular core (AR. 10:1) fiber (Fiber 4) with circular cladding and D-shaped cladding. The same launched signal powers and pumping radius are assumed as in Fig. 17.

Tables (2)

Tables Icon

Table 1 Fiber parameters used in simulation

Tables Icon

Table 2 Output powers in the investigated fibers

Equations (24)

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

E s (x,y,t)= P 01 E 01 (x,y)+ P 11 E 11 (x,y) e iΔωt
E s (x,y,z,t) z = i 2 k c 2 E s (x,y,z,t) i[ k 2 (x,y,z,t) k c 2 ] 2 k c E s (x,y,z,t)+g(x,y,z,t) E s (x,y,z,t)
k(x,y,z,t)= ω c n core (x,y,z,t) c
g(x,y,z,t)= 1 2 [ σ s a +( σ s a + σ s e ) n u (x,y,z,t)] N Yb (x,y)
n core (x,y,z,t)= n core + dn dT ΔT(x,y,z,t)
n u (x,y,z,t)= P p (z,t) σ p a h ν p A p + I s (x,y,z,t) σ s a h ν s P p (z,t)( σ p a + σ p e ) h ν p A p + I s (x,y,z,t)( σ s a + σ s e ) h ν s + 1 τ
E s (x,y,z,t) z = i 2 k c 2 E s (x,y,z,t)
E s (x,y,z,t)= 1 2π E s ( k x , k y ,z,t) e i k x x e i k y y d k x d k y
E s ( k x , k y ,z,t) z =i k x 2 2 k c E s ( k x , k y ,z,t)+i k y 2 2 k c E s ( k x , k y ,z,t)
ϕ( k x , k y )= Δz 2 ( k x 2 + k y 2 ) 2 k c
E s (x,y,z,t) z = i[ k 2 (x,y,z,t) k c 2 ] 2 k c E s (x,y,z,t)+g(x,y,z,t) E s (x,y,z,t)
ϕ(x,y,t)=Δz k 2 (x,y,t) k c 2 2 k c
G(x,y,ω| x',y')= n=0 F n (y,y') P n (x,x', ω)
F n (y,y')= 1 WK sin( nπ W y)sin( nπ W y')
P n (x,x',ω)={ exp[ σ n (2H| xx' | )]exp[ σ n (2Hxx')] +exp[ σ n | xx' |] exp[ σ n (x+x')] }/ σ n (1exp[2 σ n H])
σ n 2 = ( nπ W ) 2 +iω ρC K
ρC ΔT(x,y,z,t) t =Q(x,y,z,t)+K( 2 ΔT(x,y,z,t) x 2 + 2 ΔT(x,y,z,t) y 2 )
Q(x,y,z,t)= N Yb (x,y)[ ν p ν s ν p ][ σ p a ( σ p a + σ p e ) n u (x,y,z,t)] P p (z,t) A p
ΔT(x,y,t)=Real[ m=0 1 x',y' q(x',y', ω m ) G(x,y, ω m | x',y' ) exp(i ω m t)]
q(x',y', ω m )=ΔxΔy i=0 N t 1 Q(x',y', t i )exp(i ω m t i )
d P p (z,t) dz = P p (z,t) A p [( σ p a + σ p e ) n u (x,y,z,t) σ p a ] N Yb (x,y)dxdy
P 01 (z,t)= | E s (x,y,z,t) E 01 (x,y)dxdy [ E 01 (x,y)] 2 dxdy | 2 P 11 (z,t)= | E s (x,y,z,t) E 11 (x,y)dxdy [ E 11 (x,y)] 2 dxdy | 2
χ 1,2 (Δω)= dn dT k 2 K β 1,2 Im[A(Δω)](1 λ s λ p )
Im(A(Δω))= Ω E 01 (x,y) E 11 (x,y)× ( Ω d Im(G(x,y,x',y',Δω)) E 01 (x',y') E 11 (x',y') dx'dy')dxdy
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.