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

Influence of the third energy level on the gain dynamics of EDFAs: analytical model and experimental validation

Open Access Open Access

Abstract

We report an analytical model and experimental validation of the temporal dynamics of 3-level system fiber amplifiers. The model predictions show a good agreement with the measured pump power to output power and the pump power to output phase transfer functions in an EDFA pumped at 976 nm, as well as with the typical literature values for the spontaneous lifetime of the involved energy levels. The measurements show a linear relation between the effective lifetime of the meta-stable level and the output power, and a filtering of the temperature-induced phase-shift due to the quantum defect at a sufficiently high frequency modulation.

© 2016 Optical Society of America

1. Introduction

Although the current interferometric gravitational wave detectors (GWDs) operate with laser sources emitting at a wavelength of 1064 nm, the next generation will most probably utilize cryogenically cooled silicon mirrors to reduce the thermal noise. That change would demand laser sources at 1.5 μm [1, 2], since silicon is not transparent at 1 μm but at wavelength above 1.4 μm. Moreover, there is a clear tendency to increase the optical power in order to reduce the quantum noise and increase the sensitivity at high frequencies. This makes Er3+-doped fiber amplifiers (EDFAs) promising candidates for the next generation of GWDs. In this context, a robust knowledge of temporal and gain dynamics of the amplifier is required to design proper frequency and power stabilization systems which are able to work in a relatively wide range of frequencies. Pumping Er3+ at 1480 nm permits modeling the amplifier as a 2-level system, where the excited energy level decays to the ground energy level producing optical emission at 1.5 μm. Such systems have been widely described [3, 4]. However, because of the low power of commercially available single-mode lasers at 1480 nm, it is common to utilize pump lasers at 976 nm [5]. This not only reduces the quantum efficiency due to a higher quantum defect, but also implies the fast relaxation transition from the upper state level to the metastable intermediate level, which modifies the amplifier’s temporal dynamics. Nevertheless, these systems are usually modeled as 2-level systems as well [6], neglecting this fast relaxation process and leading to discrepancies and behaviors in the temporal dynamics that a 2-level model cannot predict. On the other hand, an extensively used configuration are Er3+:Yb3+co-doped fibers to take advantage of the higher absorption of Yb3+around 980 nm. In this kind of amplifiers, the upper energy level of Yb3+is excited, from which the energy is then transferred to the upper state level of Er3+ [7]. This solution also involves fast relaxation processes which must be modeled and, in any case, it is not possible to find a full analytical solution for the corresponding transfer function [8]. In purely Er3+-doped amplifiers, the pump power modulation couples to the amplified output by the gain, since the modulation of the population density of the intermediate energy level leads to modulation of emission at 1.5 μm. The pump power modulation is coupled to the output phase via temperature produced due to the high quantum defect of Er3+ pumped around 980 nm. Additionally, the pump power modulation can couple to the output phase by Kramer-Kronig-Relation (KKR) which can become noticeable when pump power is too low to produce a relevant population inversion [9,10].

In this paper we report an analytical model to describe the temporal dynamics of a 3-level system by means of its pump to amplified output transfer functions. We carried out two different experiments. In the first experiment we measured the pump to signal transfer function of an EDFA up to 150 kHz as higher frequencies are not relevant for GWDs. We compare these transfer functions with the model’s prediction showing a good agreement with both the magnitude and phase of the amplified signal. In the second experiment we simulated and measured the transfer function of the pump power to the phase of the output signal, which suggests a low-pass behavior of the upper energy level, responsible for the fiber heating by multiphonon emission when the transition to the intermediate energy level occurs, i.e. the quantum defect.

2. Analytical model

The relevant energy levels and transitions of Er3+ doped silica when it is pumped at 976 nm are shown in Fig. 1. At first, the energy level 4I11/2 is excited, increasing its population density. This level then decays to the level 4I13/2 in a non-radiative process which releases phonons, i.e. heat. After that, a radiative transition towards the ground energy level 4I15/2 takes place emitting photons around 1.5 μm.

 figure: Fig. 1

Fig. 1 Energy level diagram of Er3+doped in silica. ni represents the population densities of the energy levels, Wji are the rates between energy levels, and τmp and τ are the spontaneous lifetimes of the 4I11/2 and 4I13/2 levels respectively.

Download Full Size | PDF

Considering this energy level diagram, the corresponding rate equations for the population densities of the energy levels in an EDFA are

0=pn1n2n3
n2t=W12n1W21n2n2τ+n3τmp
n3t=W13n1W31n3n3τmp
where p is the doping concentration (in ions per m3) and with rates
W13=ΓpσpabsPp(z,t)AcW31=ΓpσpemPp(z,t)Ac
W12=ΓsσsabsPs(z,t)AcW21=ΓsσsemPs(z,t)Ac.
Pp(z, t) and Ps(z, t) are the pump and seed power levels (in units of photons per second) at the point z and the time t. σsabs, σpabs, σsem and σpem are the absorption and emission cross sections at the pump and the seed wavelength. In addition, Γs and Γp are the overlap factors of the seed and the pump light with the core area Ac of the fiber. These overlap factors can be expressed as
Γp,s=1exp(2rc2wp,s2)
where rc is the core diameter and
wp,src(0.65+1.619Vp,s1.5+2.879Vp,s6)
is either the mode radius of the pump or seed light (Vp,s is the V-parameter of the fiber at the pump or seed wavelength) [11]. The evolution of Pp(z, t) and Ps(z, t) along the fiber at any time t is given by
Ppz=Γp(n3σpemn1σpabs)Pp=Ac(W31n3W13n1)
Psz=Γs(n2σsemn1σsabs)Ps=Ac(W21n2W12n1).

Substituting Eq. (8) in Eq. (3) and Eq. (9) in Eq. (2) and applying the variation of the Bononi-Rusch model reported by Novak and Moesle [12] by integrating along the fiber leads to

N2t=Ps(z=0)Ps(z=L)N2τ+N3τmp
N3t=Pp(z=0)Pp(z=L)N3τmp
where N2=Ac0Ln2dz and N3=Ac0Ln3dz. Eqs. (10) and (11) can be further simplified by using the following relations for the seed and pump power at z = L with the seed and pump power at z = 0
Pp(z=L)=Pp(Z=0)eBpN3Cp+rpσpabsAcN2
Ps(z=L)=Ps(Z=0)eBsN2Cs+rsσsabsAcN3
with
Bp=ΓpAc(σpabs+σpem)Cp=Γpσpabsp
Bs=ΓsAc(σsabs+σsem)Cs=Γsσsabsp.

Substituting Eqs. (12) and (13) in Eqs. (10) and (11) yields

N2t=Ps(z=0)(1exp(BsN2Cs+ΓsσsabsAcN3))N2τ+N3τmp
N3t=Pp(z=0)(1exp(BpN3Cp+ΓpσpabsAcN2))N3τmp.

Assuming now a small modulation (mp≪1) of the input pump power

Pp(z=0,t)=Pp,in(1+mpeiωt)
and corresponding small (δ2, δ3≪1) modulations of the population densities N2(t) and N3(t) around their steady solutions N20 and N30
N2(t)=N20(1+δ2ei(ωt+ϕ2))
N3(t)=N30(1+δ3ei(ωt+ϕ3))
we can substitute Eqs. (18)(20) in Eq. (17), leading to
N3t=Pp,in(1+mpeiωt)(1e(BpN30Cp+rpσpabsAcN20)eBpN30δ3exp(i(ωt+ϕ3))erpσpabsAcN20δ2exp(i(ωt+ϕ3)))N30τmpN30τmpδ3e(i(ωt+ϕ3)).

Since δ2 and δ3 are small, we can approximate eδ21+δ2 and eδ31+δ3. Furthermore, we can neglect higher order terms (e.g. δ22 or δ2δ3). Using Eq. (12) finally gives

N3tPp0(z=0)Pp0(z=L)N30τmp=0since it is steady state+(Pp0(z=0)Pp0(z=L))mpeiωtPp0(z=L)BpN30δ3ei(ωt+ϕ3)+Pp0(z=L)ΓpσpabsAcN20δ2ei(ωt+ϕ2)N30τmpδ3ei(ωt+ϕ3).

Using Eq. (20) and multiplying Eq. (22) with eiωtmp yields

N30δ3mpeiϕ3=Pp0(z=0)Pp0(z=L)(1ΓpσpabsAcN20δ2mpeiϕ2)(iω+BpPp0(z=L)+1τmp).

In general, the term Nk0δkmpeiϕk represents the complex transfer function of the population density in the level k since Nk0δk is the magnitude of the population density modulation around its steady state, mp is the input modulation and eiϕk is the phase factor. Therefore, as can be seen in Eq. (23), the transfer function N30δ3mpeiϕ3 of the upper level 4I11/2 depends on the transfer function N20δ2mpeiϕ2 of the meta-stable level 4I13/2. If we now assume full pump power absorption, as desired in practical fiber amplifiers, we can use Pp0(z=L)0, and the transfer function N30δ3mpeiϕ3 of the upper level becomes

N30δ3mpeiϕ3=Pp0(z=0)(1τmp+iω)
which is a simple low-pass with a cutoff frequency 1τmp. Analogously, the transfer function of the meta-stable level 4I13/2 can also be obtained analytically and is given by
N20δ2mpeiϕ2=1τmpPs0(z=L)ΓsσsabsAcBsPs0(z=L)+1τ+iωN30δ3mpeiϕ3.

As mentioned, under the assumption of a complete pump light absorption, N30δ3mpeiϕ3 is a single low-pass with cutoff frequency 1τmp and thus, the transfer function N20δ2mpeiϕ2 of the meta-stable level is a double low-pass with cutoff frequencies at 1τmp and BsPs0(z=L)+1τ.

In order to obtain the transfer function of the amplified signal, we use Eq. (13) and assume again that the population densities N2(t) and N3(t) and the seed power Pp(z = L) are modulated around their steady state solutions such that

N2(t)=N20(1+δ2ei(ωt+ϕ2))
N3(t)=N30(1+δ3ei(ωt+ϕ3))
Ps(z=L)=Ps0(z=L)(1+msei(ωt+ϕs)).

Substitution of Eqs. (26)(28) in Eq. (13), using again the approximation eδ ≈ 1 + δ … (δ ≪1) and neglecting higher order terms (e.g. δ22 or δ2δ3) yields

Ps0(z=L)(1+msei(ωt+ϕs))=Ps0(z=L)(1+BsN20δ2ei(ωt+ϕ2)+ΓsσsabsAcN30δ3ei(ωt+ϕ3))
which can be simplified to
msei(ωt+ϕs)=BsN20δ2ei(ωt+ϕ2)+ΓsσsabsAcN30δ3ei(ωt+ϕ3).

Multiplication with eiωtmp leads to

msmpeiϕs=BsN20δ2mpeiϕ2+ΓsσsabsAcN30δ3mpeiϕ3
where the left hand side (msmpeiϕs) is the transfer function of the amplified signal, and the transfer functions N20δ2mpeiϕ2 and N30δ3mpeiϕ3 are known (see Eqs. (24) and (25)). Substituting Eqs. (24) and (25) in Eq. (31) yields
msmpeiϕs=Pp0(z=0)ΓsσsabsAc1τmpω2+iω(BsAcτmpΓsσsabs+1τω3+iωBsPs0(z=L)+1τω1+iω)
which can be written as
msmpeiϕs=Kω2+iω(ω3+iωω1+iω)
where K, ω1, ω2 and ω3 are given by
K=Pp0(z=0)ΓsσsabsAc
ω1=BsPs0(z=L)+1τ
ω2=1τmp
ω3=BsAcτmpΓsσsabs+1τ.

Thus, the transfer function of the amplified seed is a low-pass 1ω2+iω multiplied with the term ω3+iωω1+iω that is either a damped low-pass if ω3 > ω1 or a damped high-pass if ω1 > ω3. Comparing Eqs. (35) and (37), it can be established a threshold of the amplified output power level, in watts, to define this feature as

Ps0(z=L)ω1>ω3ω3>ω1AcτmpΓsσsabshcλs.

Where h is the Planck constant and λs is the seed wavelength. For example, if we consider a fiber with core diameter of 8.5 μm, σsabs=1.18×1025m2, τmp = 9 μs [13] and Γs≈1, then ω3 > ω1 is satisfied while Ps0(z=L)<6.75W, otherwise ω1 > ω3.

It must be noticed that Eq. (32) appears as a double low-pass if ω3 is sufficiently large and ω3 can also be expressed as function of the absorption and emission cross sections by using the definition of Bs shown in Eq. (15)

ω3=1τmpσsabs+σsemσsabs+1τ.

3. Pump-to-output signal transfer function

To validate the model, in particular Eq. (32), we compared it to the measured pump power to output power transfer function of an Er3+-doped fiber amplifier pumped at 976 nm up to 150 kHz. In this section we present the setup, results and fit of the analytical model.

3.1. Setup

The experimental setup is shown in the Fig. 2. It consisted of a commercial 14-meters long single-mode Er3+-doped silica fiber with core diameter of 8.5 μm and cladding diameter of 125 μm manufactured by CorActive. The nominal absorption at 976 nm was 1 dB per meter. The active fiber was pumped by a FBG-stabilized single-mode laser at 976 nm with a maximum output power of 600 mW and seeded at 1572 nm by a DFB diode with ~MHz linewidth. The current of the pump laser was modulated with the sweep signal from a commercial network analyzer, producing the subsequent optical modulation. The modulation depth was always kept at 10% of the mean pump power and the input seed power was constant at 65 mW. A 976/1572 nm WDM was spliced to the output of the active fiber to eliminate any possible residual pump power. Additionally, a band-pass filter centered at 1570 nm with a 12 nm bandwidth was used. The signal was detected by a photodiode with a bandwidth of 1.2 GHz. A sample of the pump power was taken as reference signal from the 1572/976 nm WDM just before the active fiber and detected by another photodiode with a 1.2 GHz bandwidth. At the maximum pump power, the output signal was 250 mW.

 figure: Fig. 2

Fig. 2 Setup used to measure the pump to signal transfer functions of an EDFA. WDM: Wavelength division multiplexer, PD: Photodiode, BPF: Band-pass filter.

Download Full Size | PDF

3.2. Results

The measured pump power to output power transfer functions are shown in Fig. 3. The magnitude shows a low-pass behavior with its roll-off frequency increasing with the output power. Furthermore, all curves reach a slope of −40 dB/dec at frequencies above ~10 kHz. Besides this, the phase presents a maximum delay of approximately −150°. In Fig. 4, the model developed in Section 2, i.e. Eq. (32), is fitted to the transfer function measured at a pump power of 482 mW and 250 mW output power. It shows a good agreement with the magnitude and the phase in the whole frequency range. Note that the frequency axis was extended in order to show how the model predicts the magnitude and phase at higher frequencies. The limiting factor for measuring higher frequencies, and eventually observe the predicted damped low-pass in the magnitude, was the sensitivity of the network analyzer.

 figure: Fig. 3

Fig. 3 Pump to amplified signal transfer functions for different pump and output power levels. Top: magnitude. Bottom: phase. The legend shows the pump power followed by the output signal power. The magnitude has been calculated as 20log10(VinVref), where Vin and Vref are the voltage signal delivered by the output photodiode, and the voltage signal delivered by the reference photodiode, respectively (see Fig. 2).

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Pump power to output power transfer function for a pump power of 482 mW and an output power of 250 mW together with the fitted model. The frequency axis is deliberately extended to show the model prediction up to 1 MHz.

Download Full Size | PDF

As stated in Section 2 (see Eqs. (32)(37)), the relations between the cutoff frequencies and the spontaneous lifetimes are given by

ω1=BsPs0(z=L)+1τ
ω2=1τmp
where ω1 and ω2 are the inverse of the effective lifetimes of the energy levels 4I13/2 and 4I11/2 respectively and τ and τmp are their corresponding spontaneous lifetimes. The analytical model predicts that ω1 increases linearly with output power with a minimum value of 1/τ. On the other hand ω2 is constant with value 1/τmp. These predictions are in good agreement with the values obtained from the measured transfer functions. Table. 1 shows the values of 1/ω1 and 1/ω2 resulting from the model fitting to the transfer functions for each output power. The mean lifetime of the energy level 4I11/2 resulted to be 7.9 μs with a standard deviation of 1.007 μs. This deviation can be attributed to uncertainties in the measurements and fitting processes as well as other physical mechanisms that are commented further in Section 4.2. The obtained value of 7.9 μs is in agreement with the typical values in the literature, which are usually around 9∼10 μs for Er3+-doped silica fibers [13]. Furthermore, the relation between the spontaneous lifetime of the meta-stable level τ, the cutoff frequency ω1/2π of its pump power to output power transfer function, and the output amplified power, is already noticeable in Fig. 3 as a shift in the cutoff frequencies with power that extends the range where the magnitude is not filtered. To calculate the spontaneous lifetime τ, one can linearly fit the evolution of the cutoff frequency ω1 to the output power and extend the linear fit to an output power of zero to obtain the value of 1/τ as shown in Fig. 5. In our case, the calculated value was τ = 8.69 ms. Given the low number of samples to perform the fit (5), and the noticeable difference between some of them and the linear fit, we can assume that there is a not negligible error in the final result. Nevertheless, the coefficient of determination was R2 > 0.9 and τ = 8.69 ms is in agreement with common values in the literature that are usually around 10 ms for Er3+-doped silica fibers [14].

Tables Icon

Table 1. Effective lifetimes 1/ω1 and 1/ω2 for each measured output signal power level, corresponding to the energy levels 4I13/2 and 4I11/2 respectively.

 figure: Fig. 5

Fig. 5 Evolution of ω1 with output power and a corresponding linear fit. The inset shows the extension to an output power of zero such that ω1 = 1/τ.

Download Full Size | PDF

4. Pump-to-output phase-shift transfer function

An effective technique to increase the optical power of laser sources that can be potentially used in the future GWDs is the coherent beam combination (CBC) of multiple fiber amplifiers. It has been demonstrated that Er3+-doped and Yb3+-doped fibers can be used as phase actuators by pump power modulation [15–17], avoiding the disadvantages of piezo-electrical or electro-optical actuators. In such configuration it becomes critical to stabilize the phase of the fiber amplifiers, for which a good understanding of the pump-power to optical phase-shift transfer function is required.

In this section we present both simulations and experimental measurements of the pumppower to output phase-shift transfer function in an EDFA. The results are in agreement with the expected phase shift due to temperature, which is the dominant mechanism, and also show consistency with the analytical model in Eq. (32) and measurements presented beforehand in Section 3.2.

4.1. Setup

In Fig. 6 the setup designed to measure the pump power to output phase-shift transfer function of the amplifier is presented. It consisted of a Mach-Zender interferometer with the EDFA described in Section 3.1 in one of its arms. The sweep signal from the network analyzer was used to modulate the pump laser power, thus producing a modulation of the population density in the energy level 4I11/2 of the Er + ions. This leads to a modulation of the phonon emission due to the quantum defect. Therefore, the temperature generated in the fiber is modulated as well, producing slight changes in the refractive index of the core via the thermo-optical effect that eventually change the optical path difference between the interferometer arms. A probe signal passing through both arms of the interferometer can be used to determine the phase-shift by measuring the power of the resulting interference at the output. In the setup shown in Fig. 6 the pump and seed lasers are the same ones used in Section 3.1. To measure the relative phase-shift, a laser diode at 1310 nm was added to provide the probe signal which was introduced to both arms of the interferometer by a 50:50 coupler. It was a DFB laser with a ∼MHz linewidth delivering an optical power of 5 mW. In the passive arm, a 40 m-long fiber stretcher was used as an actuator to stabilize the interferometer and a polarization controller was installed to improve the contrast of the interference. In order to reduce excessive coupling of the frequency noise of the 1310 nm laser to the power at the output of the interferometer, a piece of approximately 26 m of standard single-mode passive fiber (SMF-28) was attached after the EDFA. The interferometer was closed by another 50:50 coupler and the resulting interference signal was filtered by a bandpass filter centered at 1310 nm with a 12 nm bandwidth. Again, two broadband photodiodes with 1.2 GHz bandwidth were used to sense the interference and reference signals from the interferometer output and the EDFA input pump power respectively. A commercial feedback controller Digilock-110 from Toptica Photonics was used to stabilize the interferometer output power by compensating the optical phase acting with a control signal on the stretcher. In the range of frequencies in which the stabilization loop had gain to lock the interferometer, the pump power to phase-shift transfer function was measured by means of measuring the control signal acting on the stretcher. This signal is provided together with the reference signal to the network analyzer to obtain the pump power to phase-shift transfer function. At frequencies at which the stabilization loop did not have enough gain to lock the interferometer, the error signal measured by the photodiode at the output of the interferometer was used instead of the control signal.

 figure: Fig. 6

Fig. 6 Setup used to measure the pump power to phase-shift transfer function. FPGA: Digilock field programmable gate array. WDM: Wavelength division multiplexer. SMF-28: Passive single-mode fiber. BPF: Bandpass filter.

Download Full Size | PDF

4.2. Simulation and results

Using a simulation tool previously developed in our group [18], we performed simulations of the temperature-induced phase-shift in the EDFA. The simulation assumes the generation of heat in the core and calculates its radial propagation in the fiber depending on its geometry. However, it does not take into account the physical origin of the heat generation. Furthermore, the simulated transfer function excludes any temperature-independent effect that may occur in the EDFA such as KKR. In Fig. 7, both the measured and simulated phase-shift magnitudes are shown in a frequency range from 100 mHz to 80 kHz. The maximum measurable frequency was limited by the sensing photodiode and network analyzer sensitivity. It can be seen how both phase-shift magnitudes rapidly fall at frequencies between 100 mHz and 3 Hz and then keep falling slower until 10 kHz. Within this range the simulation predicts fairly well this tendency. This shows that the phase-shift is dominated by temperature and filtered by the fiber geometry and thermal properties. Note that the simulations have demonstrated a strong dependency of the frequency response on the fiber geometry at certain frequency ranges, specially with the core and cladding diameters, thermal conductivity, heat transfer coefficient [18] and, similarly, their possible small variations along the fiber. Thus, the accurate knowledge of these parameters is particularly important to achieve an accurate reproduction of the phase-shift dynamics. Nonetheless, a discrepancy starting around 20 kHz between the measured and simulated transfer functions can be noted, which could not be reproduced in the simulations using realistic parameters within the margins guaranteed by the fiber specifications. To study this further, we performed new measurements and simulations with higher frequency resolution to calculate the equivalent transfer function of this discrepancy as shown in Fig. 8. Since this means that the temperature generation reacts to a lesser extent to pump power modulation than in the simulation, we can assume this effect is a consequence of the transfer function of the population density modulation of the energy level 4I11/2, which governs the phonon emission. Therefore it should follow Eq. (24), which corresponds to a single low-pass function. By fitting a low-pass function to the corresponding discrepancy between the simulated and measured transfer functions, the resulting cutoff frequency is f ≈ 16.8 kHz, leading to an effective lifetime of τmp=12πf9.4μs, which is very close to the τmp obtained in the experiment in Section 3.2 and typical values on the literature for the lifetime of the upper state level 4I11/2 [13]. Thus, we can state that the temperature-induced phase-shift of the EDFA is filtered by the transfer function of the energy level 4I11/2.

 figure: Fig. 7

Fig. 7 Pump power modulation induced phase-shift, i.e. the magnitude of the corresponding transfer function. The curves have been scaled in order to overlap and the absolute magnitudes are arbitrary.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 Detail of the discrepancy between the phase-shift transfer functions.

Download Full Size | PDF

Nevertheless, it is worth to note that the fitted low-pass function and the mentioned discrepancy start differing slightly at approximately 80 kHz. This gap is a mere 5 dB/dec and there are several factors that could explain it. The most immediate is the inexact knowledge of the fiber properties and their slight variations along it. In this context, the small variations of the core size are specially influential [18]. Besides this, at high frequencies it is also possible that refractive index changes due to the Kramers-Kronig relations [19] and transitions between Stark splits in the energy levels 4I11/2 and 4I13/2 become relevant, since the temperature-induced phase-shift gets less dominant as the frequency increases, as stated before. None of these features are included in the simulations and further investigations are needed to fully understand their influence. Moreover, these additional mechanisms are difficult to characterize by means of the transfer function because fitting and simulation inaccuracies cannot be screened. However, the small gap shown in Fig. 8 between the calculated discrepancy and the single low-pass function can be an early evidence of their effects.

5. Conclusions

We have presented an analytical model of the temporal dynamics of EDFAs pumped at 976 nm and experimentally tested it by means of the pump power to output power and pump power to output phase-shift transfer functions. Even though we experimentally demonstrated the suitability of the developed model with an Er3+-doped fiber amplifier, it can also be applied to any 3-level system, e.g. thulium [20]. We found good agreements with the measurements in the range from 100 mHz to 150 kHz and with the typical values of the lifetimes of the energy levels used in the literature. At sufficiently high frequencies and assuming full pump power absorption, the pump power to output power transfer function is a low-pass multiplied by a damped low-pass. The phase of this transfer function does not reach −180°, a fact that simplifies the design of stabilization loops at such frequencies. The first cutoff frequency of the pump power to output power transfer function, and hence the effective lifetime τ is given by the transfer function of the ion population density of the energy level 4I13/2 and depends on its spontaneous lifetime and the output power, whereas the second cutoff frequency, which remains constant, is caused by the transfer function of the population density of the energy level 4I11/2. The modulation of the population density in the energy level 4I11/2 produces a modulation of the phonon emission rate, leading to a heat generation modulation. The pump power then couples to the output phase by the thermo-optical effect. It has been shown that the transfer function of the population density in the energy level 4I11/2, described in the analytical model, can also predict the filtering of the output phase-shift induced by temperature.

Funding

European Commission’s Seventh Framework Program (FP7-PEOPLE-2013-ITN) (606176)

References and links

1. ET Science Team, Einstein gravitational wave Telescope Conceptual Design Study (2011).

2. P. Punturo, “The third generation of gravitational wave observatories and their science reach,” Class. Quantum Grav. 27, 084007 (2010). [CrossRef]  

3. A. Saleh, R. Jopson, J. Evankow, and J. Aspell, “Modeling of gain in Erbium-doped fiber amplifiers,” IEEE Photon. Tech. Lett. 2(10), 714–717 (1990). [CrossRef]  

4. E. Desurvire, M. Zirngibl, H. Presby, and D. DiGiovanni, “Dynamic gain compensation in saturated Erbium-doped fiber amplifiers,” IEEE Photon. Tech. Lett. 3(5), 453–455 (1991). [CrossRef]  

5. V. Kuhn, D. Kracht, J. Neumann, and P. Wessels, “Er-doped photonic crystal fiber amplifier with 70 W of output power,” Opt. Lett. 36(16), 3030–3032 (2011). [CrossRef]   [PubMed]  

6. B. Pedersen, “Small-signal Erbium-doped fibre amplifiers pumped at 980 nm: a design study,” Opt. and Quantum Electron. 26(3), 273–284 (1994). [CrossRef]  

7. H. Dong, L. Sun, and C. Yan, “Energy transfer in lanthanide upconversion studies for extended optical applications,” Chem. Soc. Rev. 44(6), 1608–1634 (2015). [CrossRef]  

8. M. Steinke, J. Neumann, D. Kracht, and P. Wessels, “Gain dynamics in Er3+:Yb3+co-doped fiber amplifiers,” Opt. Express 23, 14946–14959 (2015). [CrossRef]   [PubMed]  

9. J. Martin and J. Mollier, “Experimental characterization of phase-shift induced by complex susceptibility,” Proc. SPIE 5460, 169 (2004). [CrossRef]  

10. H. Tünnermann, D. Kracht, J. Neumann, and P. Wessels, “Gain dynamics and refractive index changes in fiber amplifiers: a frequency domain approach,” Opt. Express 20(12), 13539–13550 (2012). [CrossRef]   [PubMed]  

11. G. P. Agrawal, Nonlinear Fiber Optics (Academic Press, 2007).

12. S. Novak and A. Moesle, “Analytical Model for gain modulation in EDFAs,” J. Lightwave Technol. 20(6), 975 (2002). [CrossRef]  

13. C. Layne, W. Lowdermilk, and M. Weber, “Multiphonon relaxation of rare-earth ions in oxide glasses,” Phys. Rev. B 16(1), 10–20 (1977). [CrossRef]  

14. F. Sanchez, P. Le Boudec, P. François, and G. Stephan, “Effects of ion pairs on the dynamics of Erbium-doped fiber lasers,” Phys. Rev. A 48, 2220 (1993). [CrossRef]   [PubMed]  

15. H. Tünnermann, J. Neumann, D. Kracht, and P. Wessels, “All-fiber phase actuator based on an erbium-doped fiber amplifier for coherent beam combining at 1064 nm,” Opt. Lett. 36(4), 448–450 (2011). [CrossRef]   [PubMed]  

16. H. Tünnermann, Y. Feng, J. Neumann, D. Kracht, and P. Wessels, “All-fiber coherent beam combining with phase stabilization via differential pump power control,” Opt. Lett. 37(7), 1202–1204 (2012). [CrossRef]   [PubMed]  

17. A. Fotiadi, N. Zakharov, O. Antipov, and P. Mégret, “All-fiber coherent combining of Er-doped amplifiers through refractive index control in Yb-doped fibers,” Opt. Lett. 34(22), 3574–3576 (2009). [CrossRef]   [PubMed]  

18. H. Tünnermann, J. Neumann, D. Kracht, and P. Wessels, “Frequency resolved analysis of thermally induced refractive index changes in fiber amplifiers,” Opt. Lett. 37(17), 3597–3599 (2012). [CrossRef]   [PubMed]  

19. R.N. Liu, I.A. Kostko, R. Kashyap, K. Wu, and P. Kiiveri, “Inband-pumped, broadband bleaching of absorption and refractive index changes in Erbium-doped fiber,” Opt. Commun. 255(1), 65–71 (2005). [CrossRef]  

20. T. Komukai, T. Yamamoto, T. Sugawa, and Y. Miyajima, “Upconversion pumped thulium-doped fluoride fiber amplifier and laser operating at 1.47 μ m,” IEEE J. Quantum Electron. 31(11), 1880–1889 (1995). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (8)

Fig. 1
Fig. 1 Energy level diagram of Er3+doped in silica. ni represents the population densities of the energy levels, Wji are the rates between energy levels, and τmp and τ are the spontaneous lifetimes of the 4I11/2 and 4I13/2 levels respectively.
Fig. 2
Fig. 2 Setup used to measure the pump to signal transfer functions of an EDFA. WDM: Wavelength division multiplexer, PD: Photodiode, BPF: Band-pass filter.
Fig. 3
Fig. 3 Pump to amplified signal transfer functions for different pump and output power levels. Top: magnitude. Bottom: phase. The legend shows the pump power followed by the output signal power. The magnitude has been calculated as 20 log 10 ( V in V ref ), where Vin and Vref are the voltage signal delivered by the output photodiode, and the voltage signal delivered by the reference photodiode, respectively (see Fig. 2).
Fig. 4
Fig. 4 Pump power to output power transfer function for a pump power of 482 mW and an output power of 250 mW together with the fitted model. The frequency axis is deliberately extended to show the model prediction up to 1 MHz.
Fig. 5
Fig. 5 Evolution of ω1 with output power and a corresponding linear fit. The inset shows the extension to an output power of zero such that ω1 = 1/τ.
Fig. 6
Fig. 6 Setup used to measure the pump power to phase-shift transfer function. FPGA: Digilock field programmable gate array. WDM: Wavelength division multiplexer. SMF-28: Passive single-mode fiber. BPF: Bandpass filter.
Fig. 7
Fig. 7 Pump power modulation induced phase-shift, i.e. the magnitude of the corresponding transfer function. The curves have been scaled in order to overlap and the absolute magnitudes are arbitrary.
Fig. 8
Fig. 8 Detail of the discrepancy between the phase-shift transfer functions.

Tables (1)

Tables Icon

Table 1 Effective lifetimes 1/ω1 and 1/ω2 for each measured output signal power level, corresponding to the energy levels 4I13/2 and 4I11/2 respectively.

Equations (41)

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

0 = p n 1 n 2 n 3
n 2 t = W 12 n 1 W 21 n 2 n 2 τ + n 3 τ mp
n 3 t = W 13 n 1 W 31 n 3 n 3 τ mp
W 13 = Γ p σ p abs P p ( z , t ) A c W 31 = Γ p σ p em P p ( z , t ) A c
W 12 = Γ s σ s abs P s ( z , t ) A c W 21 = Γ s σ s em P s ( z , t ) A c .
Γ p,s = 1 exp ( 2 r c 2 w p,s 2 )
w p,s r c ( 0.65 + 1.619 V p,s 1.5 + 2.879 V p,s 6 )
P p z = Γ p ( n 3 σ p em n 1 σ p abs ) P p = A c ( W 31 n 3 W 13 n 1 )
P s z = Γ s ( n 2 σ s em n 1 σ s abs ) P s = A c ( W 21 n 2 W 12 n 1 ) .
N 2 t = P s ( z = 0 ) P s ( z = L ) N 2 τ + N 3 τ mp
N 3 t = P p ( z = 0 ) P p ( z = L ) N 3 τ mp
P p ( z = L ) = P p ( Z = 0 ) e B p N 3 C p + r p σ p abs A c N 2
P s ( z = L ) = P s ( Z = 0 ) e B s N 2 C s + r s σ s abs A c N 3
B p = Γ p A c ( σ p abs + σ p em ) C p = Γ p σ p abs p
B s = Γ s A c ( σ s abs + σ s em ) C s = Γ s σ s abs p .
N 2 t = P s ( z = 0 ) ( 1 exp ( B s N 2 C s + Γ s σ s abs A c N 3 ) ) N 2 τ + N 3 τ mp
N 3 t = P p ( z = 0 ) ( 1 exp ( B p N 3 C p + Γ p σ p abs A c N 2 ) ) N 3 τ mp .
P p ( z = 0 , t ) = P p,in ( 1 + m p e i ω t )
N 2 ( t ) = N 2 0 ( 1 + δ 2 e i ( ω t + ϕ 2 ) )
N 3 ( t ) = N 3 0 ( 1 + δ 3 e i ( ω t + ϕ 3 ) )
N 3 t = P p,in ( 1 + m p e i ω t ) ( 1 e ( B p N 3 0 C p + r p σ p abs A c N 2 0 ) e B p N 3 0 δ 3 exp ( i ( ω t + ϕ 3 ) ) e r p σ p abs A c N 2 0 δ 2 exp ( i ( ω t + ϕ 3 ) ) ) N 3 0 τ mp N 3 0 τ mp δ 3 e ( i ( ω t + ϕ 3 ) ) .
N 3 t P p 0 ( z = 0 ) P p 0 ( z = L ) N 3 0 τ mp = 0 since it is steady state + ( P p 0 ( z = 0 ) P p 0 ( z = L ) ) m p e i ω t P p 0 ( z = L ) B p N 3 0 δ 3 e i ( ω t + ϕ 3 ) + P p 0 ( z = L ) Γ p σ p abs A c N 2 0 δ 2 e i ( ω t + ϕ 2 ) N 3 0 τ mp δ 3 e i ( ω t + ϕ 3 ) .
N 3 0 δ 3 m p e i ϕ 3 = P p 0 ( z = 0 ) P p 0 ( z = L ) ( 1 Γ p σ p abs A c N 2 0 δ 2 m p e i ϕ 2 ) ( i ω + B p P p 0 ( z = L ) + 1 τ mp ) .
N 3 0 δ 3 m p e i ϕ 3 = P p 0 ( z = 0 ) ( 1 τ mp + i ω )
N 2 0 δ 2 m p e i ϕ 2 = 1 τ mp P s 0 ( z = L ) Γ s σ s abs A c B s P s 0 ( z = L ) + 1 τ + i ω N 3 0 δ 3 m p e i ϕ 3 .
N 2 ( t ) = N 2 0 ( 1 + δ 2 e i ( ω t + ϕ 2 ) )
N 3 ( t ) = N 3 0 ( 1 + δ 3 e i ( ω t + ϕ 3 ) )
P s ( z = L ) = P s 0 ( z = L ) ( 1 + m s e i ( ω t + ϕ s ) ) .
P s 0 ( z = L ) ( 1 + m s e i ( ω t + ϕ s ) ) = P s 0 ( z = L ) ( 1 + B s N 2 0 δ 2 e i ( ω t + ϕ 2 ) + Γ s σ s abs A c N 3 0 δ 3 e i ( ω t + ϕ 3 ) )
m s e i ( ω t + ϕ s ) = B s N 2 0 δ 2 e i ( ω t + ϕ 2 ) + Γ s σ s abs A c N 3 0 δ 3 e i ( ω t + ϕ 3 ) .
m s m p e i ϕ s = B s N 2 0 δ 2 m p e i ϕ 2 + Γ s σ s abs A c N 3 0 δ 3 m p e i ϕ 3
m s m p e i ϕ s = P p 0 ( z = 0 ) Γ s σ s abs A c 1 τ mp ω 2 + i ω ( B s A c τ mp Γ s σ s abs + 1 τ ω 3 + i ω B s P s 0 ( z = L ) + 1 τ ω 1 + i ω )
m s m p e i ϕ s = K ω 2 + i ω ( ω 3 + i ω ω 1 + i ω )
K = P p 0 ( z = 0 ) Γ s σ s abs A c
ω 1 = B s P s 0 ( z = L ) + 1 τ
ω 2 = 1 τ mp
ω 3 = B s A c τ mp Γ s σ s abs + 1 τ .
P s 0 ( z = L ) ω 1 > ω 3 ω 3 > ω 1 A c τ mp Γ s σ s abs h c λ s .
ω 3 = 1 τ mp σ s abs + σ s em σ s abs + 1 τ .
ω 1 = B s P s 0 ( z = L ) + 1 τ
ω 2 = 1 τ mp
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.