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

Relationship between glass structure and spectroscopic properties in Er3+/Yb3+/Al3+/P5+-doped silica glasses

Open Access Open Access

Abstract

In this study, Er3+/Yb3+/Al3+/P5+-doped silica (simplified as EYAPS) glasses with different P/Al ratios ranging from zero to infinity were fabricated via the sol-gel method combined with high-temperature sintering. The absorption, emission and fluorescence lifetime of Yb3+ and Er3+ ions as well as the energy transfer efficiency from Yb3+ to Er3+ ions were recorded. The composition-dependent macroscopic properties were correlated to the glass structures, and probed by pulse electron paramagnetic resonance (EPR) and Raman spectroscopy. Results show that the spectral properties of Er3+ and Yb3+ ions and their local environment as well as the global glass network structure of EYAPS glasses are strongly dominated by the P/Al ratio. With the increase of the P/Al ratio, pulse EPR shows that rare earth ions gradually moved from a silicon and aluminum rich environment to a phosphorus rich environment. Raman spectroscopy shows that the maximum phonon energy of EYAPS samples gradually increases from 1200 to 1326 cm−1 due to the formation of AlPO4-like units and P = O double bonds. These structural changes lead to a gradual increase of peak absorption and emission cross sections of Er3+ ions at 1.5 µm, as well as the energy transfer efficiency of 2F5/2 level of Yb3+ to 4I11/2 level of Er3+ ions.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

High-power eye-safe erbium-doped silica fiber (EDF) laser operated at 1.5µm is attractive for its wide application in the fields of free-space telecommunications, remote sensing, and range finding [14]. So far, the maximum output power (∼656 W) from a single EDF has lagged far behind that of ytterbium doped silica fiber (> 10 KW) [5,6]. There are two key factors that limit the power scaling of EDF [5,7,8]: small absorption cross section at 980 nm and serious concentration quenching effect.

Laser diodes (LDs) with wavelengths of 980, 1480 and 1530 nm are usually used as pump sources of EDF [1]. Compared with 1480 and 1530 nm LDs, the 980 nm LD has higher output power and electrical-to-optical efficiency as well as lower price. However, the absorption cross section of Er3+ ion at 980 nm is one order of magnitude smaller than that of Yb3+ ion [1]. In order to increase the absorption coefficient of EDF at 980 nm and suppress the nonlinear effect of EDF due to the long fiber length, Yb3+ ions are always co-doped into EDF to sensitize Er3+ ions. The Yb3+ ions can absorb most of the pump light at 980 nm and transfer absorbed energy to the adjacent Er3+ ions via cross relaxation. Furthermore, co-doping with Yb3+ ions can reduce Er3+ clusters and inhibit concentration quenching effect [9].

Early studies show that the rare earth (RE) ions tend to form clusters and micro-phase separation even at a low doping levels (∼1000 ppm by weight), due to their poor solubility in pure silica glass [7,8,1012]. Such RE clustering gives rise to drastically reduce fluorescence lifetime of RE ions, significantly increase laser threshold and substantially reduce slope efficiency of RE-doped silica fibers. This phenomenon is called concentration quenching effect [8]. Co-doping with Al2O3 or P2O5 can effectively improve the solubility of RE ions and prevent clustering of RE ions in silica glass [1214]. On the other hand, in order to improve the energy transfer efficiency from Yb3+ to Er3+, high concentrations of phosphorus are always doped into silica matrix. However, single-doping with large amounts of phosphorus (>10 mol% P2O5) has following disadvantages, such as a detrimental central dip in the refractive index profile, a high core numerical aperture (NA) and a high core background loss [15]. Early studies [16,17] show that Al and P tend to bond with each other in Al2O3-P2O5-SiO2 glass, and its global network structure is strongly dominated by the excessed Al2O3 or P2O5 component. Co-doping with equimolar Al2O3 and P2O5 can decrease the refractive index of fiber core due to the formation of AlPO4-like unit [17,18]. In addition, co-doping with Al and P can effectively disperse RE ion clusters and inhibit the central refractive index dip of fiber core [1921].

The spectral properties of Er3+ ions are closely correlated to their local environment. The spectral properties and local environment of Er3+ ions in different host glasses (such as silicate, aluminate, phosphate and fluoride) have been systematically studied using optical absorption and photoluminescence (PL) as well as x-ray absorption fine structure (XAFS) methods [2224]. Effect of Al2O3 content on the spectral properties and local environment of Er3+ ions in Er3+/Al3+ co-doped silica glasses have been also systematically studied using PL and XAFS methods [2527]. The PL spectral shape and peak cross section (σem) of Er3+ ions at 1.53 µm in Er3+/Al3+, Er3+/Ge4+, Er3+/P5+ doped silica glasses were comparatively studied by Wang et al. [28], the σem of the studied glasses follows the sequence of Er3+/Ge4+ > Er3+/Al3+ > Er3+/P5+. Using pulse EPR spectroscopy, Saitoh et al. [29] confirmed that Er3+ ions preferentially coordinate to phosphorus rather than aluminum in Er3+/Al3+ or Er3+/P5+-doped silica glasses. However, the effects of P and Al co-doping ratio (P/Al ratio) on the spectral properties and local environment of Er3+ ions as well as the energy transfer mechanism between Yb3+ and Er3+ ions in Er3+/Yb3+/Al3+/P5+ co-doped silica glasses are not fully understood.

In this work, Er3+/Yb3+/Al3+/P5+ co-doped silica glasses with different P/Al ratio were fabricated by sol-gel method combined with high temperature sintering. The doping concentrations of Yb3+ and Er3+ ions were fixed. The effects of P/Al ratio on the spectral properties of Yb3+ and Er3+ ions and energy transfer efficiency between them were systematically studied. The local environment of RE ions was probed using pulse EPR. The global glass network structure was studied by Raman spectroscopy. Based on the above experiments, the relationships between glass structure and spectroscopic properties in Er3+/Yb3+/Al3+/P5+ co-doped silica glasses were established. Due to the influence of fiber drawing process, there will be a little difference between the performance of optical fiber and that of fiber preform [30,31]. However, to some extent, this work can provide theoretical guidance for the optimization of fiber core composition.

2. Experimental details

In this study, Er3+/Yb3+/Al3+/P5+ (denoted as EYAPS) doped silica glasses were fabricated using the sol-gel method combined with high temperature sintering. TEOS, ErCl3·6H2O, YbCl3·6H2O, AlCl3·6H2O, H3PO4, and C2H5OH were used as precursors. Pure water was used to sustain the hydrolysis reaction. All precursors were mixed and stirred at 25°C to form homogeneous doped gel. Then the sol was heated from 30 to 1000 under oxygen atmosphere to decompose the hydroxyl and organics. The obtained dry sol powder was then melted at 1650-1750 °C for 2 h in vacuum state to obtain transparent bulk glass. The details of the fabrication process are given in Ref. [32]. In order to calculate the energy transfer efficiency from 2F5/2 level of Yb3+ to 4I11/2 level of Er3+ ions, the corresponding Er-free Yb3+/Al3+/P5+ doped silica (denoted as YAPS) glass samples were prepared using the same method.

Table 1 shows mean theoretical compositions of EYAPS and YAPS series glass samples. The doping concentrations of Er2O3 and Yb2O3 were fixed at 0.1 mol% and 0.5 mol%, respectively. P/Al mole ratio increases from zero to infinity. Inductively coupled plasma atomic emission spectrometer (ICP-AES) test indicated that the actual Yb2O3, Er2O3 and Al2O3 contents are very close to their theoretical values, while the actual P2O5 content is slightly lower than its theoretical content due to the volatilization of phosphorus during high temperature sintering.

Tables Icon

Table 1. Mean compositions of EYAPS and YAPS glass samples (in mol%)

Bulk glasses were cut and polished to small chips with a thickness of 2 mm and a diameter of about 15 mm for the spectroscopic tests. Powder samples with a weight of about 200 mg were used for EPR tests.

The optical absorption spectra were tested via a Lambda 950 UV–Vis–NIR spectrophotometer. The PL spectra excited at 896 nm for Yb3+ and Er3+ ions with Xe lamp, and the fluorescence decay curves of Er3+: 4I13/2 and Yb3+: 2F5/2 energy levels pumped with a pulsed 980 nm laser diode (LD) were recorded on a high-resolution Edinburgh Instruments, FLS 920 spectrofluorometer. The testing step lengths of absorption and PL spectra are 1 nm, which are much smaller than the peak widths of absorption and PL spectra. The Raman measurement using a 488 nm argon ion laser as exciting source was carried out by a Renishaw InVia Raman Microscope.

Pulse EPR experiments were performed using an X-band EPR spectrometer (E-580 BRUKER ELEXSYS) at the temperature of 4 K. In order to probe the effect of P/Al ratio on the local environment of RE ions, the two-pulsed (π/2-τ-π-τ-echo) echo detected field-swept (EDFS) and four-pulsed (π/2-τ-π/2-T1-π-T2-π/2-τ-echo) two-dimensional hyperfine sublevel correlation (2D-HYSCORE) EPR spectra were recorded. The π/2, π, and τ values in these two experiments corresponded to 16, 32, and 136 ns, respectively.

3. Results and discussion

3.1 Effect of P/Al ratio on the spectral properties

Figure 1 shows the absorption spectra of EYAPS series samples in the ultraviolet and visible (UV-VIS) region. The UV absorption edges show a blue shift and the UV absorption intensities show a decrease with increasing P/Al ratio. These changes are mainly correlated to the presence of Yb3+ ions, because the content of Yb3+ ions (1mol%) are much higher than that of Er3+ ions (0.2mol%) and the matrix glass has not absorption in UV region. According to early studies [3335], the strong UV absorption in the range of 190–290 nm were primarily due to the charge-transfer (CT) transition Yb3+ ions and the position of CT bands shifted to a shorter wavelength with increasing electronegativity of the next nearest neighbor atoms (Al/Si/P) of Yb3+ ions. With increasing P/Al, Yb3+ ions gradually moved from silicon rich and aluminum rich environment to phosphorus rich environment, that is, Yb-O-Al and Yb-O-Si linkages are gradually replaced by Yb-O-P linkage (See Fig. 5). The absorption band at 330 nm is primarily due to the 4f-5d transition of Yb2+ ions. With the increase of P/Al ratio, the absorption intensity of Yb2+ ion decreases. Since the absorption of Yb2+ ion covers the range of 190-600nm [33], the absolute absorption intensities of 4G11/2 (379 nm) and 4F11/2 (520 nm) levels of Er3+ ions decrease with the decrease of Yb2+ ion absorption intensity. When P/Al > 1, the absorption peak of Yb2+ is almost invisible as shown in EYAPS2 and EYAPS3 samples. It indicates that the reduction of Yb3+ to Yb2+ can be effectively inhibited when P/Al > 1. Similar results have been reported in our previous studies in Yb3+/Al3+/P5+ doped silica glasses [18,34].

 figure: Fig. 1.

Fig. 1. UV-VIS absorption spectra of EYAPS series samples, the inset photographs show the large-sized and transparent EYAPS0, EYAPS1, EYAPS2 and EYAPS3 samples, respectively.

Download Full Size | PDF

The inset photographs in Fig. 1 show the large-sized and transparent EYAPS0, EYAPS1, EYAPS2 and EYAPS3 samples, respectively. With increasing P/Al ratio, the color of glass gradually changes from light yellow to light pink. Light yellow is mainly caused by Yb2+ absorption, while light pink is mainly caused by Er3+ absorption. The change of glass color also indicates that Yb2+ ions were inhibited gradually with the increase of P/Al ratio.

Figure 2(a) shows the near-infrared (NIR) absorption spectra of EYAPS series samples near 1 µm, showing a typical 2F7/22F5/2 absorption peak of Yb3+ ions. Because of low doping concentration and low absorption cross section at 980 nm of Er3+ ions, the absorption peak intensities of Er3+ ions in this band are very weak, which is completely covered by the absorption peak of Yb3+ ions. The absorption spectra of Yb3+ ions show two different types of line-shape in Fig. 2(a). When P/Al < 1, the absorption peak widths of EYAPS0 and EYAPS1 samples at 915 nm are relatively fuller. When P/Al > 1, the absorption peak intensities of EYAPS2 and EYAPS3 samples at 915 nm are obviously weakened, and the peak width is narrowed. The absorption peak intensities at 975 nm are also decreased in EYAPS2 and EYAPS3 samples which have P/Al ratio larger than 1. Similar result has been reported in our earlier work [18]. Therefore, the effect of P/Al ratio on spectral properties of Yb3+ ions will not be discussed in detail below.

 figure: Fig. 2.

Fig. 2. NIR absorption spectra of EYAPS series samples near 1 µm (a) and 1.5 µm (b).

Download Full Size | PDF

Figure 2(b) shows the NIR absorption spectra of EYAPS series samples near 1.5 µm, showing a typical 4I15/24I13/2 absorption peak of Er3+ ions. The absorption peaks also show two different types of line-shape in case of P/Al < 1 and P/Al > 1. Compared with P/Al < 1 samples, the absorption peak intensity of P/Al > 1 samples at 1480 nm decreases significantly, and the strongest absorption peak near 1530 nm appears redshifted. When P/Al < 1, the strongest absorption peak locates at 1528 nm and its full width at half maximum (FWHM) is about 40 nm; when P/Al > 1, the strongest absorption peak is at 1532 nm and its FWHM is about 18 nm (See Table 2 for more detail).

Tables Icon

Table 2. The peak absorption (λ abs) and emission (λ em) wavelength, full width at half maximum (FWHM) of absorption and emission bands (∼ 1.53 µm), and peak absorption (σabs) and emission (σem) cross sections for the transition between 4I13/2 and 4I15/2 of Er3+ ions with different P/Al ratio.

Figure 3(a) shows the PL spectra of glass samples excited by 896 nm Xe lamp. All PL spectra are normalized to the same intensity at 1.5 µm. The PL bands at 1 µm and 1.5 µm are due to the 2F5/22F7/2 of Yb3+ and 4I13/24I15/2 of Er3+ ions, respectively. With the increase of P/Al ratio, the PL intensity at 1 µm decreases. It suggests that high P/Al ratio will reduce the PL intensity of Yb3+ ions. Compared with P/Al < 1 samples, the widths of 1.5 µm PL band in P/Al > 1 samples are obviously narrowed. At the same time, the strongest PL peak appears redshifted. For P/Al < 1 samples, the strongest PL peak is located at 1530 nm, and the FWHM of 1.5 µm PL band is about 46 nm; For P/Al > 1 samples, the strongest PL peak is redshifted to 1535 nm, and the FWHM of 1.5 µm PL band is narrowed to 21 nm (See Table 2 for more detail). The narrow FWHM (∼21 nm) in P/Al > 1 sample implies that the Er-doped silica fiber with P/Al > 1 is not suitable for broadband amplification in C-band (1530 ∼ 1565 nm). However, the PL band above 1580 nm is relatively flat in P/Al > 1 samples, which is conducive to improve the gain flatness of Er3+-doped fiber in L-band (1565-1625 nm).

 figure: Fig. 3.

Fig. 3. (a) Normalized PL spectra, and (b) Emission intensity ratio of Er3+ ions at 1530 and Yb3+ ions at 975 nm (I1530/I975) of EYAPS series samples.

Download Full Size | PDF

Figure 3(b) shows the ratio of the strongest PL intensity (1530 nm) of Er3+:4I13/24I15/2 to that (975 nm) of Yb3+:2F5/22F7/2 (denoted as I1530/I975). Increasing P/Al ratio from zero to infinity, the I1530/I975 ratio increases from four to fourteen. It implies that increasing P/Al ratio can improve the energy transfer efficiency from 2F5/2 level of Yb3+ to 4I11/2 level of Er3+ ions.

Table 2 lists the peak wavelength and full width at half maximum (FWHM) of absorption and emission bands (∼ 1.53 µm) for the transition between 4I13/2 and 4I15/2 of Er3+ ions in EYAPS series glasses. The peak absorption (σabs) and emission (σem) cross sections of Er3+ ions at about 1.53 µm are also added into Table 2. σabs and σem are calculated using the Beer–Lambert law and McCumber theory respectively as follows [36]:

$${\sigma _{abs}} = \frac{{2.303}}{{{N_0}L}}OD(\lambda )$$
$${\sigma _{emi}} = {\sigma _{abs}}\frac{{{Z_l}}}{{{Z_u}}} exp \left[ {\frac{{hc}}{{kT}}\left( {\frac{1}{{{\lambda_0}}} - \frac{1}{\lambda }} \right)} \right]$$

Where, OD(λ) is optical density obtained directly from the absorption tests, N0 is the Er3+ concentration (ions/cm3) determined by ICP and density measurements, and L is the glass sample thickness. ZU and ZL are the partition functions of the upper and lower energy levels of Er3+ ions, respectively. The ratio of ZL and ZU (ZL/ZU) is approximately 8/7 at room temperature for Er3+ ions. The letters h, c, K, T represent the Planck constant, light speed, Boltzmann constant and Kelvin temperature, respectively. λ0 is the wavelength of the zero-line energy. Here, λ0 is taken as the wavelength corresponding to the intersection of absorption and emission bands at 1.5 µm.

Table 3 presents the fluorescence lifetime of Yb3+:2F5/2 level (${\tau _{Yb}}$) and Er3+:4I13/2 level (${\tau _{Er}}$) in Er-containing EYAPS series samples, as well as the fluorescence lifetime of Yb3+:2F5/2 level ($\tau _{Yb}^0$) in the corresponding Er-free YAPS series samples. The energy transfer efficiency (η) from Yb3+:2F5/2 level to Er3+:4I13/2 level is also added into Table 3.

Tables Icon

Table 3. Lifetime of Er3+:4I13/2 level (${{\boldsymbol {\tau} }_{{\boldsymbol {Er}}}}$) and Yb3+:2F5/2 level (${{\boldsymbol {\tau} }_{{\boldsymbol {Yb}}}}$) in EYAPS series samples; Lifetime of Yb3+:2F5/2 level (${\boldsymbol {\tau} }_{{\boldsymbol {Yb}}}^0$) in Er3+-free YAPS series samples; Energy transfer coefficients (${\bf{\mathrm{\eta}} }$) from Yb3+:2F5/2 level to Er3+:4I13/2 level

The η value can be calculated using Eq. (3):

$${\eta } = 1 - \frac{{{\tau _{Yb}}}}{{\tau _{Yb}^0}}$$

With the increase of P/Al ratio, the lifetime of Er3+:4I13/2 level (${\tau _{Er}}$) in EYAPS samples shows a slight decline from 11.98 to 10.6 ms. The lifetime of Yb3+:2F5/2 level (${\tau _{Yb}}$) in Er-containing EYAPS series samples has no obvious change. It is within the allowable range of measurement uncertainties (≤ 10µs). The lifetime of Yb3+:2F5/2 level ($\tau _{Yb}^0$) in Er-free YAPS series samples shows a significant increase from 1253 to 2011 µs. Compared with YAPS series samples, the shorter lifetime of Yb3+:2F5/2 level (130∼180µs) in EYAPS series samples implies a strong energy transfer from Yb3+:2F5/2 level to Er3+:4I13/2 level. With the increase of P/Al ratio, the energy transfer efficiency increases from 86.3% (in EYAPS0) to 92.1% (in EYAPS3) as shown in Table 3. It indicates that increasing P/Al ratio in ETAPS series glasses promotes the forward energy transfer from Yb3+:2F5/2 level to Er3+:4I13/2 level. At the same time, it inhibits the reverse energy transfer from Er3+:4I13/2 level to Yb3+:2F5/2 level.

3.2 Effect of P/Al ratio on the glass structure

Figure 4 shows EDFS spectra. The EDFS line-shape and its strongest peak position are related to the site heterogeneity and g-anisotropy of RE3+ ions [33,37]. All EDFS spectra consist of broad asymmetric curves, which are correlated with the inherent topological disorder of glass. With the increasing of P/Al ratio, the strongest EDFS peak shifts to a lower magnetic field. It indicates that the local environment of RE3+ ions has changed in EYPAS series samples.

 figure: Fig. 4.

Fig. 4. EDFS spectra of EYPAS series samples.

Download Full Size | PDF

In order to probe the local environment of RE3+ ions, 2D-HYSCORE spectra were recorded as shown in Fig. 5. At magnetic field of 350 mT, three patterns located at 6.0, 3.9 and 3.0 MHz correspond to the Larmor frequencies of the nuclides 31P (natural abundance ∼ 100%), 27Al (natural abundance ∼ 100%), and 29Si (natural abundance ∼ 4.68%), respectively. With increasing P/Al ratio, the magnitude of 31P pattern increases gradually, but the magnitude of 29Si pattern decreases until it disappears. The magnitude of 27Al pattern increases first, and then it decreases until disappears. This result is well consistent with the early studies in Yb3+/Al3+/P5+-doped silica glasses [16,35,38]. It suggests that the introduction of a small number of Er2O3 (∼0.1mol%) does not significantly change the local environment of RE ions (Yb3+ and Er3+). In fact, the atomic radius and charge of Er3+ and Yb3+ ions are very close, indicating that they are chemically equivalent to each other. No significant difference can be detected in the HYSCORE tests for chemically equivalent paramagnetic ions. For P/Al < 1 samples, the RE3+ ions are mainly located in silicon rich and aluminum rich environment. For P/Al > 1 samples, the RE3+ ions are mainly located in phosphorus rich environment. Only phosphorus pattern is detected in EYAPS3 sample. It indicates that almost all RE3+ ions in EYAPS3 sample are surrounded by phosphorus solvent shell. While a strong 31P pattern and a weak 27Al pattern can be observed in EYAPS2 sample. It suggests that most of RE3+ ions are in phosphorus rich environment, but a small part of RE3+ ions are still located in aluminum rich environment in EYAPS2 sample. The change of the coordination environment of RE3+ ions is responsible for the change of EDFS EPR spectra as shown in Fig. 4. Oxygen atoms that coordinate to the first shell of RE ions are not detected, owing to the low natural abundance (∼0.038%) of magnetic nucleus 17O.

 figure: Fig. 5.

Fig. 5. 2D-HYSCORE spectra of EYAPS0 (a), EYAPS1 (b), EYAPS2 (c) and EYAPS3 (d) samples.

Download Full Size | PDF

In order to study the effect of P/Al ratio on global glass network structure, the Raman spectra of EYAPS series samples were recorded as shown in Fig. 6. The Raman spectrum of undoped silica glass (α-SiO2) is also added into Fig. 6 for comparison. A broad and strong band at 440 cm−1 as well as five weak bands peaked at 490, 606, 800, 1160 and 1200 cm−1 can be observed in α-SiO2. All these Raman peaks in α-SiO2 also can be observed in EYAPS series samples. Among them, the 490 and 600 cm−1 bands are ascribed to the planar four-fold (D1) ring and planar three-fold (D2) ring, respectively [39]. The 440, 800, 1060 and 1200 cm−1 bands are ascribed to symmetric bending (ω1), symmetric stretching (ω3), TO and LO asymmetric stretching (TO-ω4 and LO-ω4) vibration of Si-O-Si bonds, respectively [40,41]. The Raman spectra of P/Al < 1 samples are very similar to that of α-SiO2, the maximum phonon energy of P/Al < 1 samples is about 1200 cm−1, and it is primarily from the LO-ω4 vibration modes of Si-O-Si bonds. Compared with Raman spectra of P/Al < 1 samples, a narrow and sharp peak peaked at 1326 cm−1 can be observed in P/Al > 1 samples, this peak provides maximum phonon energy for P/Al > 1 samples, and it is due to the stretching vibration of P = O bond in P(3) unit (Its structural model is O = P-O3/2) [17,40]. In addition, a broad and strong band at 1000-1250 cm−1 can only be observed in EYAP2 sample, but not in EYAP3 sample. This broad band peaked at 1145 cm−1 is usually seen as the evidence of the existence of AlPO4-like units [19]. Sample EYAP0 is only co-doped with aluminum and but not phosphorus, while EYAP3 is only co-doped with phosphorus and but not aluminum. No Al-O-P linkages were formed in EYAP0 and EYAP3 samples, so no AlPO4-like units are observed in the Raman spectra of these two samples. For EYAP1 sample, the signal of AlPO4-like units is very weak owing to its low doping amount of phosphorus.

 figure: Fig. 6.

Fig. 6. Raman spectra of pure silica (α-SiO2) and EYAPS series glasses.

Download Full Size | PDF

3.3 Correlation between spectral properties and glass structure

As described in Section 3.1, the spectral properties of Er3+ and Yb3+ ions and the energy transfer efficiency between them in EYAPS series samples, are highly dependent on the P/Al ratio. They should be correlated to the structures of these glasses as revealed in Section 3.2.

From Tables 2 and 3, increasing P/Al ratio, the peak wavelength of absorption and emission bands of Er3+ ions at 1.53 µm is redshifted, FWHM is narrowed, and the peak absorption and emission cross sections are increased. At the same time, the energy transfer efficiency of 2F5/2 level of Yb3+ to 4I11/2 level of Er3+ ions gradually increase. The changes of these spectral parameters are correlated to the changes of local environment of Er3+ ions and phonon energy of glass matrix.

The spectral broadening of RE ions includes homogeneous broadening and inhomogeneous broadening. Among them, the homogeneous broadening is correlated to the interactions between RE and RE as well as RE and glass matrix. The inhomogeneous broadening primarily originates from the heterogeneity of RE sites [1,42]. In silicate glasses, the contribution of inhomogeneous broadening is greater than that of homogeneous broadening [42]. For P/Al < 1 samples, pulsed EPR (See Fig. 5) shows that the RE3+ ions are mainly surrounded by Si and Al atoms. Among them, Si mainly exists in the form of SiO4/2, and Al consists of three structural states (four, five and six coordinated aluminum) as evidenced by nuclear magnetic resonance (NMR) [16,17]. The various coordination environments of RE ions lead to the larger inhomogeneous broadening of absorption and PL bands of Er3+ ions at 1.5µm. While for P/Al > 1 samples, pulsed EPR (See Fig. 5) shows that the RE3+ ions are mainly surrounded by P atoms. When P/Al > 1, excessive P mainly exists in the form of P(3) structure (Its structural model is O = P-O3/2) as evidenced by Raman (See Fig. 6) and NMR [16,17]. The monotonous coordination environment of RE ions leads to the narrower absorption and PL bands of Er3+ ions at 1.5µm. Generally speaking, the absorption and emission cross sections of Er3+ ions at 1.5 µm are inversely proportional to their FWHMs. Therefore, increasing P/Al ratio causes the FWHMs of Er3+ ions at 1.5 µm bands become narrow, while both absorption and emission cross sections increase (See Table 2).

Figure 7 shows the schematic diagram of forward and reverse energy transfer from Yb3+:2F5/2 level to Er3+:4I11/2 level. Early studies show that the lifetimes of Er3+:4I11/2 in germanate, aluminosilicate, and phosphate glasses are about 15, 4, and 0.1 µs, respectively [43,44]. While the maximum phonon energy of these three types of glasses increases gradually, which are 820, 1200, 1326 cm−1, respectively. This result suggests that the maximum phonon energy of glass matrix has a great impact on the lifetime of Er3+: 4I11/2. The high phonon energy of glass matrix will speed up the non-radiation transition rate of Er3+ ions from 4I11/2 level to 4I13/2 level. Therefore, it decreases the fluorescent lifetime of Er3+:4I11/2 level, and reduces the number of Er3+ ion population in 4I11/2 level. As a result, it inhibits the reverse energy transfer from Er3+:4I11/2 level to Yb3+:2F5/2 level, and improves the forward energy transfer from Yb3+:2F5/2 level to Er3+:4I11/2 level.

 figure: Fig. 7.

Fig. 7. Schematic diagram of energy transfer between Yb3+:2F5/2 level and Er3+:4I11/2 level.

Download Full Size | PDF

Raman spectra show that the maximum phonon energy of EYAPS series samples gradually increases from 1200 to 1326 cm−1 (See Fig. 6). The forward energy transfer efficiency from Yb3+:2F5/2 level to Er3+:4I11/2 level in EYAPS series samples increases gradually from 86.3% (in EYAPS0) to 92.1% (in EYAPS3) with increasing P/Al ratio (See Table 3). When P/Al < 1, the relatively lower maximum phonon energy (∼1200 cm−1) of glass matrix decreases the non-radiative transition rate (∼1/4 µs−1) from 4I11/2 level to 4I13/2 level, and it results in the increased probability of reverse energy transfer from Er3+:4I11/2 level to Yb3+:2F5/2 level. At the same time, the forward energy transfer efficiency from Yb3+:2F5/2 level to Er3+:4I11/2 level relatively decreases. When P/Al > 1, the relatively higher maximum phonon energy (∼1326 cm−1) of glass matrix increases the non-radiative transition rate (∼1/0.1 µs−1) from 4I11/2 level to 4I13/2 level, and it leads to the decreased reverse energy transfer probability from Er3+:4I11/2 level to Yb3+:2F5/2 level. At the same time, the forward energy transfer efficiency from Yb3+:2F5/2 level to Er3+:4I11/2 level greatly increases.

4. Conclusions

In this work, Er3+/Yb3+/Al3+/P5+-doped silica glasses with different P/Al ratios ranging from zero to infinity were fabricated using the sol-gel method combined with high-temperature sintering. Combining with the optical absorption, photoluminescence (PL), fluorescence decay curves, Raman, and pulse electron paramagnetic resonance (EPR) spectroscopies, the effect of P/Al ratio on the spectroscopic properties of Er3+ and Yb3+ ions and their local environment as well as the global glass network structure were studied systematically, and the relationships between spectral properties and glass structures were established.

The results show that the spectroscopic properties of Er3+ ions as well as their local environment present two distinct states with P/Al = 1 as the boundary. When P/Al < 1, the Er3+ ions are mainly surrounded by Si and Al atoms, the absorption and fluorescence peaks of Er3+ ions at 1.5 µm are wider. When P/Al > 1, the Er3+ ions are mainly surrounded by P atoms, the absorption and fluorescence peaks of Er3+ ions at 1.5 µm get narrower. Raman spectra show that the maximum phonon energy of Er3+/Yb3+/Al3+/P5+-doped silica glasses gradually increases from 1200 to 1326 cm−1 with increasing P/Al ratio due to the formation of AlPO4-like unit and P = O double bonds. Due to these structural changes, the peak absorption and emission cross sections of Er3+ ions at 1.5 µm as well as the energy transfer efficiency of 2F5/2 level of Yb3+ to 4I11/2 level of Er3+ ions gradually increase with increasing P/Al ratio, at the expense of a slight decrease in fluorescence lifetime of Er3+ ions at 1.5 µm. Furthermore, the formation of Yb2+ ions can be effectively inhibited when P/Al > 1, which leads to the UV transmittance increasing of the glass.

This work demonstrates that the local environment and spectral properties in Er3+/Yb3+/Al3+/P5+ co-doped silica glass with P/Al ratio more than one is similar to that of Er3+/Yb3+/P5+ co-doped silica glass. In Er3+/Yb3+/P5+ co-doped fibers, the central dip of refractive index profile and high refractive index of fiber core are the common problems, which may be solved by an optimized Er3+/Yb3+/Al3+/P5+ co-doping composition.

Funding

National Natural Science Foundation of China (61775224, 61875216).

Acknowledgements

Chongyun Shao and Fan Wang contributed equally to this work. This work was financially supported by the National Natural Science Foundation of China (Grant Nos. 61775224, 61875216).

Disclosures

The authors declare no conflicts of interest.

References

1. L. V. Kotov and M. E. Likhachev, “High Power Continuous-Wave Er-doped Fiber Lasers,” in Optics, Photonics and Laser Technology 2017, P. Ribeiro, D. L. Andrews, and M. Raposo, eds. (Springer International Publishing, 2019), pp. 165–192.

2. Z. Dong, R. Xu, W. Zhang, H. Guoyu, L. Hua, J. Tian, and Y. Song, “Er-doped all-fiber laser mode-locked by graphitic carbon nitride nanosheets,” Chin. Opt. Lett. 16(8), 081401 (2018). [CrossRef]  

3. Z. Guo, Q. Hao, J. Peng, and H. Zeng, “Environmentally stable Er-fiber mode-locked pulse generation and amplification by spectrally filtered and phase-biased nonlinear amplifying long-loop mirror,” High Power Laser Sci. Eng. 7, e41 (2019). [CrossRef]  

4. J. Alaniz-Baylon, M. Durán-Sánchez, R. I. Álvarez-Tamayo, B. Posada-Ramírez, M. Bello-Jiménez, B. Ibarra-Escamilla, A. A. Castillo-Guzman, and E. A. Kuzin, “Fiber laser with simultaneous multi-wavelength Er/Yb passively Q-switched and single-wavelength Tm gain-switched operations,” Photonics Res. 7(6), 608–614 (2019). [CrossRef]  

5. H. Lin, Y. Feng, Y. Feng, P. Barua, J. K. Sahu, and J. Nilsson, “656 W Er-doped, Yb-free large-core fiber laser,” Opt. Lett. 43(13), 3080 (2018). [CrossRef]  

6. M. N. Zervas and C. A. Codemard, “High power fiber lasers: a review,” IEEE J. Sel. Top. Quantum Electron. 20(5), 219–241 (2014). [CrossRef]  

7. E. Delevaque, T. Georges, M. Monerie, P. Lamouler, and J.-F. Bayon, “Modeling of pair-induced quenching in erbium-doped silicate fibers,” IEEE Photonic. Tech. L 5(1), 73–75 (1993). [CrossRef]  

8. J. Wagener, P. Wysocki, M. Digonnet, H. Shaw, and D. DiGiovanni, “Effect of concentration on the efficiency of erbium-doped silica fiber lasers,” Proc. SPIE 1789, 80 (1993). [CrossRef]  

9. N. V. Kiritchenko, L. V. Kotov, M. A. Melkumov, M. E. Likhachev, M. M. Bubnov, M. V. Yashkov, A. Y. Laptev, and A. N. Guryanov, “Effect of ytterbium co-doping on erbium clustering in silica-doped glass,” Laser Phys. 25(2), 025102 (2015). [CrossRef]  

10. V. McGahay and M. Tomozawa, “Phase separation in rare-earth-doped SiO2 glasses,” J. Non-Cryst. Solids 159(3), 246–252 (1993). [CrossRef]  

11. R. S. Quimby, W. J. Miniscalco, and B. Thompson, “Clustering in erbium-doped silica glass fibers analyzed using 980 nm excited state absorption,” J. Appl. Phys. 76(8), 4472–4478 (1994). [CrossRef]  

12. R. S. Quimby, W. J. Miniscalco, and B. A. Thompson, “Quantitative characterization of clustering in erbium-doped silica glass fibers,” in Fiber Laser Sources and Amplifiers V (International Society for Optics and Photonics, 1994), pp. 2–12.

13. K. Arai, H. Namikawa, K. Kumata, T. Honda, Y. Ishii, and T. Handa, “Aluminum or phosphorus co-doping effects on the fluorescence and structural properties of neodymium-doped silica glass,” J. Appl. Phys. 59(10), 3430–3436 (1986). [CrossRef]  

14. V. Petit, T. Okazaki, E. H. Sekiya, R. Bacus, K. Saito, and A. J. Ikushima, “Characterization of Yb3+ clusters in silica glass preforms,” Opt. Mater. 31(2), 300–305 (2008). [CrossRef]  

15. S. Unger, A. Schwuchow, S. Jetschke, V. Reichel, A. Scheffel, and J. Kirchhof, “Optical properties of Yb-doped laser fibers in dependence on codopants and preparation conditions,” (SPIE, 2008), pp. 689016–6890111.

16. C. Shao, J. Ren, F. Wang, N. Ollier, F. Xie, X. Zhang, L. Zhang, C. Yu, and L. Hu, “Origin of Radiation-Induced Darkening in Yb3+/Al3+/P5+-Doped Silica Glasses: Effect of the P/Al Ratio,” J. Phys. Chem. B 122(10), 2809–2820 (2018). [CrossRef]  

17. B. G. Aitken, R. E. Youngman, R. R. Deshpande, and H. Eckert, “Structure−Property Relations in Mixed-Network Glasses: Multinuclear Solid State NMR Investigations of the System xAl2O3: (30-x) P2O5: 70SiO2,” J. Phys. Chem. C 113(8), 3322–3331 (2009). [CrossRef]  

18. W. Xu, J. Ren, C. Shao, X. Wang, M. Wang, L. Zhang, D. Chen, S. Wang, C. Yu, and L. Hu, “Effect of P5+ on spectroscopy and structure of Yb3+/Al3+/P5+ co-doped silica glass,” J LUMIN 167, 8–15 (2015). [CrossRef]  

19. F. Wang, C. Shao, C. Yu, S. Wang, L. Zhang, G. Gao, and L. Hu, “Effect of AlPO4 join concentration on optical properties and radiation hardening performance of Yb-doped Al2O3-P2O5-SiO2 glass,” J. Appl. Phys. 125(17), 173104 (2019). [CrossRef]  

20. M. E. Likhachev, M. M. Bubnov, K. V. Zotov, D. S. Lipatov, M. V. Yashkov, and A. N. Guryanov, “Effect of the AlPO4 join on the pump-to-signal conversion efficiency in heavily Er-doped fibers,” Opt. Lett. 34(21), 3355 (2009). [CrossRef]  

21. S. Jetschke, S. Unger, A. Schwuchow, M. Leich, and J. Kirchhof, “Efficient Yb laser fibers with low photodarkening by optimization of the core composition,” Opt. Express 16(20), 15540–15545 (2008). [CrossRef]  

22. W. J. Miniscalco, “Erbium-doped glasses for fiber amplifiers at 1500 nm,” J. Lightwave Technol. 9(2), 234–250 (1991). [CrossRef]  

23. P. M. Peters and S. N. Houde-Walter, “X-ray absorption fine structure determination of the local environment of Er3+ in glass,” Appl. Phys. Lett. 70(5), 541–543 (1997). [CrossRef]  

24. P. M. Peters and S. N. Houde-Walter, “Local structure of Er3+ in multicomponent glasses,” J. Non-Cryst. Solids 239(1-3), 162–169 (1998). [CrossRef]  

25. T. Haruna, J. Iihara, Y. Saito, K. Yamaguchi, M. Onishi, and S. Ishikawa, “Analysis of Coordination Structure around Erbium in Erbium-Doped Fiber,” SEI Tech. Review - Eng. Ed. 66, 129 (2008).

26. T. Haruna, J. Iihara, K. Yamaguchi, Y. Saito, S. Ishikawa, M. Onishi, and T. Murata, “Local structure analyses around Er3+ in Er-doped fibers with Al co-doping,” Opt. Express 14(23), 11036 (2006). [CrossRef]  

27. Q. He, F. Wang, Z. Lin, C. Shao, M. Wang, S. Wang, C. Yu, and L. Hu, “Temperature dependence of spectral and laser properties of Er3+/Al3+ co-doped aluminosilicate fiber,” Chin. Opt. Lett. 17(10), 101401 (2019). [CrossRef]  

28. Q. Wang, N. K. Dutta, and R. Ahrens, “Spectroscopic properties of Er doped silica glasses,” J. Appl. Phys. 95(8), 4025–4028 (2004). [CrossRef]  

29. A. Saitoh, S. Matsuishi, C. Se-Weon, J. Nishii, M. Oto, M. Hirano, and H. Hosono, “Elucidation of Codoping Effects on the Solubility Enhancement of Er3+ in SiO2 Glass: Striking Difference between Al and P Codoping,” J. Phys. Chem. B 110(15), 7617–7620 (2006). [CrossRef]  

30. F. Cova, M. Fasoli, F. Moretti, N. Chiodini, K. Pauwels, E. Auffray, M. T. Lucchini, E. Bourret, I. Veronese, E. D. Ippolito, and A. Vedda, “Optical properties and radiation hardness of Pr-doped sol-gel silica: Influence of fiber drawing process,” J. Lumin. 192, 661–667 (2017). [CrossRef]  

31. I. Veronese, C. De Mattia, M. Fasoli, N. Chiodini, M. C. Cantone, F. Moretti, C. Dujardin, and A. Vedda, “Role of Optical Fiber Drawing in Radioluminescence Hysteresis of Yb-Doped Silica,” J. Phys. Chem. C 119(27), 15572–15578 (2015). [CrossRef]  

32. F. Wang, L. Hu, W. Xu, M. Wang, S. Feng, J. Ren, L. Zhang, D. Chen, N. Ollier, G. Gao, C. Yu, and S. Wang, “Manipulating refractive index, homogeneity and spectroscopy of Yb3+-doped silica-core glass towards high-power large mode area photonic crystal fiber lasers,” Opt. Express 25(21), 25960–25969 (2017). [CrossRef]  

33. C. Shao, F. Xie, F. Wang, M. Guo, R. Zhu, H. Zhang, M. Wang, S. Wang, C. Yu, and L. Hu, “UV absorption bands and its relevance to local structures of ytterbium ions in Yb3+/Al3+/P5+-doped silica glasses,” J. Non-Cryst. Solids 512, 53–59 (2019). [CrossRef]  

34. S. Wang, F. Lou, C. Yu, Q. Zhou, M. Wang, S. Feng, D. Chen, L. Hu, W. Chen, M. Guzik, and G. Boulon, “Influence of Al3+ and P5+ ion contents on the valence state of Yb3+ ions and the dispersion effect of Al3+ and P5+ ions on Yb3+ ions in silica glass,” J. Mater. Chem. C 2, 4406 (2014).

35. T. Okazaki, E. H. Sekiya, and K. Saito, “P/Al atomic ratio dependence of local structure around Yb3+ ions and optical properties of Yb-Al-P-doped silica glass,” Jpn. J. Appl. Phys. 58(12), 122005 (2019). [CrossRef]  

36. D. E. McCumber, “Theory of phonon-terminated optical masers,” Phys. review 134(2A), A299–A306 (1964). [CrossRef]  

37. M. de Oliveira, T. Uesbeck, T. S. Gonçalves, C. J. Magon, P. S. Pizani, A. S. S. de Camargo, and H. Eckert, “Network Structure and Rare-Earth Ion Local Environments in Fluoride Phosphate Photonic Glasses Studied by Solid-State NMR and Electron Paramagnetic Resonance Spectroscopies,” J. Phys. Chem. C 119(43), 24574–24587 (2015). [CrossRef]  

38. T. Deschamps, N. Ollier, H. Vezin, and C. Gonnet, “Clusters dissolution of Yb3+ in codoped SiO2-Al2O3-P2O5 glass fiber and its relevance to photodarkening,” J. Chem. Phys. 136(1), 014503 (2012). [CrossRef]  

39. F. L. Galeener, “Planar rings in vitreous silica,” J. Non-Cryst. Solids 49(1-3), 53–62 (1982). [CrossRef]  

40. C. Shao, Y. Jiao, F. Lou, M. Wang, L. Zhang, S. Feng, S. Wang, D. Chen, C. Yu, and L. Hu, “Enhanced radiation resistance of ytterbium-doped silica fiber by pretreating on a fiber preform,” Opt. Mater. Express 10(2), 408 (2020). [CrossRef]  

41. F. L. Galeener and A. E. Geissberger, “Raman studies of vitreous SiO2 versus fictive temperature,” Phys. Rev. B 28(6), 3266–3271 (1983). [CrossRef]  

42. W. J. Miniscalco, “Optical and electronic properties of rare earth ions in glasses,” Opt. Eng. 71, 17–112 (2001). [CrossRef]  

43. E. F. Artem’Ev, A. G. Murzin, and Y. K. Fedorov, “Some characteristics of population inversion of the 4I13/2 level of erbium ions in ytterbium-erbium glasses,” Sov. J. Quantum Electron. 11(9), 1266–1268 (1981). [CrossRef]  

44. J. E. Townsend, W. L. Barnes, and S. G. Crubb, “Yb3+ sensitised Er3+ doped silica optical fibre with ultra high transfer efficiency and gain,” MRS Proc. 244, 143 (1991). [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 (7)

Fig. 1.
Fig. 1. UV-VIS absorption spectra of EYAPS series samples, the inset photographs show the large-sized and transparent EYAPS0, EYAPS1, EYAPS2 and EYAPS3 samples, respectively.
Fig. 2.
Fig. 2. NIR absorption spectra of EYAPS series samples near 1 µm (a) and 1.5 µm (b).
Fig. 3.
Fig. 3. (a) Normalized PL spectra, and (b) Emission intensity ratio of Er3+ ions at 1530 and Yb3+ ions at 975 nm (I1530/I975) of EYAPS series samples.
Fig. 4.
Fig. 4. EDFS spectra of EYPAS series samples.
Fig. 5.
Fig. 5. 2D-HYSCORE spectra of EYAPS0 (a), EYAPS1 (b), EYAPS2 (c) and EYAPS3 (d) samples.
Fig. 6.
Fig. 6. Raman spectra of pure silica (α-SiO2) and EYAPS series glasses.
Fig. 7.
Fig. 7. Schematic diagram of energy transfer between Yb3+:2F5/2 level and Er3+:4I11/2 level.

Tables (3)

Tables Icon

Table 1. Mean compositions of EYAPS and YAPS glass samples (in mol%)

Tables Icon

Table 2. The peak absorption (λ abs) and emission (λ em) wavelength, full width at half maximum (FWHM) of absorption and emission bands (∼ 1.53 µm), and peak absorption (σabs) and emission (σem) cross sections for the transition between 4I13/2 and 4I15/2 of Er3+ ions with different P/Al ratio.

Tables Icon

Table 3. Lifetime of Er3+:4I13/2 level ( τ E r ) and Yb3+:2F5/2 level ( τ Y b ) in EYAPS series samples; Lifetime of Yb3+:2F5/2 level ( τ Y b 0 ) in Er3+-free YAPS series samples; Energy transfer coefficients ( η ) from Yb3+:2F5/2 level to Er3+:4I13/2 level

Equations (3)

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

σ a b s = 2.303 N 0 L O D ( λ )
σ e m i = σ a b s Z l Z u e x p [ h c k T ( 1 λ 0 1 λ ) ]
η = 1 τ Y b τ Y b 0
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.