Sunteți pe pagina 1din 9

Applied Physics B (2019) 125:117

https://doi.org/10.1007/s00340-019-7233-9

Spectroscopic and waveguide properties of ­Nd3+‑doped


oxyfluorosilicate glasses
Ramachari Doddoji1,2 · G. V. Vázquez3 · R. Trejo‑Luna4 · Devarajulu Gelija5

Received: 22 February 2019 / Accepted: 27 May 2019 / Published online: 6 June 2019
© Springer-Verlag GmbH Germany, part of Springer Nature 2019

Abstract
In this work, we present the fabrication of ­Nd3+-doped oxyfluorosilicate glasses with good spectroscopic properties and
waveguiding characteristics. The amorphous nature and elemental mapping of the chemical composition for the studied
glasses were confirmed by scanning electron microscopy (SEM) and energy-dispersive spectroscopy (EDS), respectively.
Absorption, photoluminescence, and decay measurements have also been investigated. The Judd–Ofelt intensity parameters
have been evaluated from the absorption spectrum to predict the radiative properties for the emission levels ­Nd3+ ions.
The near-infrared (NIR) emission spectra recorded with 808 nm laser diode excitation show the highest emission intensity
for the 4F3/2 → 4I11/2 transition at 1.058 μm. In addition, planar waveguides were generated by either proton or carbon ion
implantation in glasses with 1.0 and 1.5 mol% N ­ d3+ concentration, showing good optical confinement from the waveguide
mode distribution and transmission properties. The obtained results show that these glasses have potential features for the
development of integrated devices, particularly the fabrication of lasers and amplifiers in a compact configuration.

1 Introduction ­(Nd3+)-doped glass materials are suitable for the develop-


ment of tunable and high-power lasers due to the wide gain
Research on special glasses based on rare-earth oxides bandwidth [5, 6]. Compared to the other R ­ E3+ [7–9], the
­( RE3+) with exceptional spectroscopic properties is of 4 4 3+
F3/2 →  I13/2 transition of ­Nd ion could be used to obtain
great importance for the development of amplifiers and efficient 1.3 μm emission even in the glass systems with
­ E3+, the neodymium
lasers [1–4]. In view of the facts on R medium–high phonon energy owing to the wide energy gap
(5500 cm−1) to the next lower level.
Among the wide glass family, silicate glass has attracted
* G. V. Vázquez
gvvazquez@cio.mx considerable interest due to its high transparency in the vis-
ible and near-infrared, high chemical stability, few defects,
Ramachari Doddoji
ramachari@tdtu.edu.vn low density, and low processing cost [10]. Thus, these
glasses are extensively used in commercial optics including
R. Trejo‑Luna
rebeca@fisica.unam.mx integrated applications [11]. The addition of ­Al2O3 modifier
in silicate glass increases the thermo-mechanical strength
Devarajulu Gelija
deva777phy@gmail.com with chemical durability [12]. Moreover, the fluorides such
as ­CaF2 and S ­ rF2 incorporated into silicate glass are good
1
Ceramics and Biomaterials Research Group, Advanced optical materials with high solubility of ­RE3+ ions and the
Institute of Materials Science, Ton Duc Thang University, phonon energy is minimized with corresponding high-effi-
Ho Chi Minh City, Vietnam
ciency luminescence. Therefore, N ­ d3+-doped oxyfluoride
2
Faculty of Applied Sciences, Ton Duc Thang University, glasses combine the merits of low-phonon energy envi-
Ho Chi Minh City, Vietnam
ronment of fluoride and desirable mechanical and chemi-
3
Centro de Investigaciones en Óptica, Loma del Bosque 115, cal properties of oxide glasses, making them attractive for
Lomas del Campestre, 37150 León, Guanajuato, Mexico
broadband optical amplifiers and lasers [12–14]. All the
4
Instituto de Física, Universidad Nacional Autónoma de aforementioned benefits offered by various chemical com-
México, Apartado Postal 20364, 01000 Mexico City, Mexico
ponents such as ­SiO2, ­Al2O3, and fluorides prompted us to
5
Department of Physics, National Chung Cheng University, prepare a stable glass system (i.e., oxyfluorosilicate glass)
Ming-Hsiung, Chia‑Yi 621, Taiwan

13
Vol.:(0123456789)
117 
Page 2 of 9 R. Doddoji et al.

to use for the formation of waveguides, which are essential Table 1  Implantation parameters used to fabricate the optical wave-
in integrated optics technology. guides
Optical waveguides play an important role for signal Sample Ion Energy Dose Inci-
propagation and the interconnections between different inte- (MeV) (ions/cm2) dence
grated optical devices. Thus, the fabrication of waveguides angle
(°)
has become more and more attractive in modern photonics
systems. Among the fabrication techniques of optical wave- 1 C2+ 5 8 × 1014 8
guides, ion implantation has been widely used because of 2 H+ 1 3 × 1016 60
its great control in waveguide parameters such as thickness
and refractive index change [15–21]. Waveguide formation
using this technique is generally based on a reduction in the focused onto a PC-controlled SP-2300i spectrograph (Acton
physical density, and hence refractive index, at the end of the Research) and detected by an InGaAs detector (Thorlabs
ion trajectory by radiation damage. Therefore, the waveguide DET10C). The fluorescence lifetime was measured using a
is confined between this low index layer—so-called “optical pulsed 808 nm laser diode. Scanning electron microscope
barrier”—and a low index material, which is usually the air (SEM) and energy-dispersive spectroscopy (EDS) maps
on the surface. were recorded using a JEOL scanning electron microscope
The purpose of this work is to synthesize oxyfluorosili- Model JSM-IT500.
cate glasses doped with different concentrations of N ­ d3+ Planar optical waveguides were fabricated on NSASCF10
ions and characterize their spectroscopic properties through (sample 1) and NSASCF15 (sample 2) glasses by ion
absorption, luminescence, and lifetime measurements. In implantation using an NEC 9SDH-2 Pelletron accelerator
addition, SEM and EDS measurements were carried out (UNAM). The implantation was simulated using the Stop-
for the compositional analysis of the glass host. In addition, ping and Range of Ions in Matter (SRIM) program [22].
we have investigated the formation of planar optical wave- Implantation parameters were selected to obtain waveguides
guides by either proton or carbon ion implantation. The with a width of a few microns (see Table 1). Images of the
waveguide characterization comprises optical microscopy, implanted zone were taken by optical microscopy from
near-field imaging, refractive index profiles, and propaga- which the damage depth was measured.
tion losses. Propagation modes were obtained by the dark mode
method using a Metricon prism coupler 2010 at a wave-
length of 633 nm [15]. From the propagation modes and
2 Materials and methods the damage depth, refractive index profiles were calculated
using a multilayer approach [23]. This method takes into
Nd 3+-doped oxyfluorosilicate glasses (NSASCF) with account both the confined modes as well as the radiation
chemical composition (mol%) of (48 − x) ­S iO 2 –10 modes; as the light travels from the prism to the substrate, it
­Al2O3–20 ­SrF2–22 ­CaF2–x ­Nd2O3 (where x = 0.1, 0.5, 1.0, can behave as a sinusoidal wave (propagation mode) or as an
and 1.5 mol% and are referred as NSASCF01, NSASCF05, evanescent wave (radiation mode), depending on the region,
NSASCF10, and NSASCF15, respectively) were prepared where it is passing.
by the conventional melt quenching technique. About 20 g Intensity mode distribution and propagation losses were
of batch composition was thoroughly crushed in an agate measured using the end-fire coupling technique [15], where
mortar and this homogeneous mixture was taken into a plati- light from an He–Ne laser operating at 633 nm was coupled
num crucible and heated in an electric furnace at 1400 °C for to the waveguides through a 10× microscope lens, and the
90 min. The melt was poured onto a preheated brass mold light was coupled out through a 20× microscope lens and
and annealed at 350 °C for 10 h to remove thermal strains. directed to either a charged coupled device (CCD) camera to
The glass samples are slowly cooled down to room tempera- acquire the near-field image, or a power detector to measure
ture and polished for optical measurements. the light intensity from the waveguide.
Refractive index (n = 1.585) measurements were real-
ized using an Abbe’s refractometer at sodium vapour lamp
wavelength (589.3 nm) for 1 mol% NSASCF10, which is 3 Results and discussion
measured for Judd–Ofelt intensity parameters. The absorp-
tion spectrum was measured using a spectrophotometer 3.1 SEM and EDS analyses
(Perkin Elmer Lambda-950) in the range of 300–1000 nm
with 1 nm resolution. The photoluminescence spectra and Figure 1 shows the SEM (inset) and EDS results of the
decay curves were recorded by exciting the glass sam- NSASCF glasses with different concentrations of N ­ d 3+
ples with an 808 nm laser diode. The emitted signal was ions. It is very clear that no nucleation parts or crystal

13
Spectroscopic and waveguide properties of ­Nd3+‑doped oxyfluorosilicate glasses Page 3 of 9  117

­ d3+ ions
Fig. 1  SEM images and EDS spectra of NSASCF glass for different concentrations of N

growth were observed in the SEM images even at the


magnification of 1 μm, revealing the formation of glasses
with high homogeneity and the absence of cracks and
unmelted portions. In addition, from these images, it is
concluded that present glasses are amorphous in nature,
because no grains can be observed. In addition, EDS ele-
mental mapping of the NSASCF glasses clearly depicts
the incorporation of different concentrations of ­Nd3+ ions
together with the presence of Si, Al, Sr, Ca, O, and F
elements. The compositional analysis in wt% of each ele-
ment of the glass host is also presented in Fig. 1. Even
in low concentration, the N­ d3+ element was detected and
the spectral data reflect values of 2.87%, 2.88%, 3.56%,
and 5.27% of ­Nd3+ that is present in the glass NSASCF01,
NSASCF05, NSASCF10, and NSASCF15, respectively.
Therefore, EDS analysis confirms the incorporation of the
­Nd3+ ions in the NSASCF glass host. Fig. 2  UV–Vis–NIR absorption spectrum of the NSASCF10 glass

13
117 
Page 4 of 9 R. Doddoji et al.

3.2 Optical absorption spectrum and Judd–Ofelt Table 2  Experimental (fexp) and calculated (fcal) oscillator strengths
parameters ­(10−6) for the 1 mol% NSASCF10 glass
Transition Wave number NSASCF10
4
The optical absorption spectrum recorded for NSASCF10 I9/2→ (cm−1)
fexp fcal
glass in the wavelength range of 300–1000  nm and the
4
assignments of absorption bands are shown in Fig. 2, as F3/2 11,376 4.65 5.18
­ d3+-doped glasses [24, 25].
illustrated in earlier studies on N 4
F5/2 + 2H9/2 12,422 14.26 13.50
4
The absorption spectrum displayed 11 inhomogeneously F7/2 + 4S3/2 13,368 16.60 17.04
broadened absorption bands due to the ­4f3 − 4f3 electronic
4
F9/2 14,641 1.02 1.30
­ d3+ ions. Among all the absorption bands,
2
transitions of N H11/2 16,000 0.48 0.36
4
the band at 805 nm is the most used for the optical pumping G5/2 + 2G−7/2 17,123 33.70 33.69
4
of ­Nd3+-based lasers, either by flash lamps or by semicon- G7/2 19,047 7.12 7.26
4
ductor GaAs laser diodes [26]. G9/2 19,493 1.95 1.18
2
For NSASCF10 glass, the experimental oscillator strength G9/2 + 2D−3/2 21,097 2.10 2.15
2
(fexp) of absorption bands is determined as [24, 25] K15/2 21,645 1.02 0.33
2
P1/2 + 2D5/2 23,201 1.01 1.30
4
fexp = 4.318 × 10−9 ∫ 𝛼(𝜐) d𝜐, (1) D3/2 + 4D1/2 28,143 9.80 10.35
δrms ± 0.46
where α(υ) is the molar absorptivity of a band at wavenum-
ber (υ) in c­ m−1. According to the Judd–Ofelt (JO) theory
[27, 28], the calculated oscillator strengths (fcal) for the Table 3  Judd–Ofelt intensity parameters (Ωλ × 10−20  cm2), and their
­ d3+:doped glass systems
trend in different N
absorption band corresponding to the electronic transition
from an initial state ψJ to a final state ψ′J′ can be estimated Glass Ω2 Ω4 Ω6 Trend
by the following equation:
NSASCF10 (present glass) 5.93 9.86 11.83 Ω6 > Ω4 > Ω2
( ) 8𝜋 2 mc𝜈 (n2 + 2)2 ∑ ‖ ‖ Nd3+:SFB [29] 8.06 4.83 11.37 Ω6 > Ω2 > Ω4
fcal 𝛹 J → 𝛹 � J � = 𝛺𝜆 (𝛹 J ‖U 𝜆 ‖𝛹 � J � )2 ,
3h(2J + 1) 9n 𝜆=2,4,6
‖ ‖ LG-750 [30] 4.60 4.80 5.60 Ω6 > Ω4 > Ω2
(2) ED-2 [31] 3.30 4.68 5.18 Ω6 > Ω4 > Ω2
where m is the mass of the electron, c is the velocity of Phosphate [32] 3.28 3.54 4.67 Ω6 > Ω4 > Ω2
light in vacuum, h is the Planck’s constant, n is refractive Fluoroborate glass [33] 4.63 2.55 6.79 Ω6 > Ω2 > Ω4
index, (n2 + 2)2/9n is the Lorentz local field correction for the Silicate [34] 4.71 4.54 5.05 Ω6 > Ω4 > Ω2
absorption band, Ωλ (λ = 2, 4, 6) are the host-dependent JO
intensity parameters, and ||Uλ|| are the doubly reduced matrix
elements of the unit tensor operator of rank λ evaluated in ­ d3+-doped glasses
is larger than those of earlier studies on N
the intermediate coupling approximation for a particular [29–34]. This indicates the higher rigidity of NSASCF10
transition and are considered to be independent on the host. glass.
The experimentally measured oscillator strengths (fexp) along
with the calculated oscillator strengths (fcal) are presented
in Table 2. The small root-mean-square deviation (δrms) of 3.3 Near‑infrared (NIR) emission and radiative
0.46 × 10−6 indicates that the deriving process between the properties
fexp and fcal oscillator strengths is reliable.
The evaluated JO intensity parameters (Ω2, 4, 6), and their Figure 3 shows the NIR emission spectra for different con-
trend in the NSASCF10 glass are compared along with some centrations of N­ d3+ in NSASCF glasses. The spectra exhibit
other reported N ­ d3+:SFB [29], LG-750 [30], ED-2 [31], one intense emission band centered at 1058 nm which cor-
phosphate [32], fluoroborate [33], and silicate [34] glasses, responds to the 4F3/2 → 4I11/2 transition and two more emis-
as shown in Table 3. The trend of the intensity parameters sion bands centered at 897 nm (4F3/2 → 4I9/2) and 1332 nm
in the present NSASCF10 glass has been found to be in (4F3/2 → 4I13/2), respectively. The radiative properties such
the order Ω6 > Ω4 > Ω2. In general, the magnitude of Ω2 as effective bandwidth (Δλeff), radiative transition probabil-
parameter depends on the site symmetry and the covalent ity (AR), experimental (βexp), and calculated (βcal) branch-
nature between rare-earth ions and ligand anions, whereas ing ratios and stimulated emission cross section (σemi) esti-
the values of Ω4 and Ω6 are related to the rigidity of the mated from the emission spectrum of NSASCF10 glass for
host medium in which the ions are situated [24, 25]. For the the 4F3/2 → 4I9/2,11/2,13/2 transitions are presented in Table 4.
present NSASCF10 glass, the Ω6 value (11.83 × 10−20 ­cm2) From the emission spectrum, the stimulated emission cross

13
Spectroscopic and waveguide properties of ­Nd3+‑doped oxyfluorosilicate glasses Page 5 of 9  117

(1.63 × 10−25  cm3), SPbKNLFNd10 (1.49 × 10−25  cm3),


TZN10 (1.32 × 10−25 cm3), and LTTNd (1.94 × 10−25 cm3)
glasses [24, 36–38].

3.4 Decay time measurements

Figure 4 shows the decay profiles of 4F3/2 level by monitoring


the emission at 1.058 μm for the different concentrations of
­Nd3+ in NSASCF glasses. For 0.1 mol% of ­Nd3+ ions concen-
tration, the decay curve exhibits single exponential nature and
turns into non-exponential at higher N ­ d3+ ion concentrations
(≥ 0.5 mol %). The lifetimes of the.
4
F3/2 level has been determined by finding the first e-folding
and average lifetime (τavg) for the single and non-exponential
decay curves, respectively [39]. The average lifetime has been
determined using the following formula:
Fig. 3  NIR-emission spectra for different ­Nd3+ ion concentrations of
∫ tI(t)dt
NSASCF glasses 𝜏avg = , (4)
∫ I(t)dt

section (σemi) of the 4F3/2 → 4I11/2 transition has been evalu-


ated using the following equation:

𝜆4p (4 )
𝜎emi = AR F3∕2 ⟶4 I11∕2 , (3)
8𝜋cn2 Δ𝜆eff

where Δλeff is the effective emission bandwidth determined


by dividing the area of the emission band by its average
height.
As shown in Table  4, the highest branching ratio (βR)
value is an attractive feature for lower threshold and higher
gain of lasers [35, 36]. The large radiative transition prob-
ability (7014 s−1) is related to a high refractive index glass,
which results in a short radiative lifetime (τrad) of 143 µs.
From the stimulated emission cross-section (6.58 × 10−20 cm2)
and gain bandwidth parameter (2.30 × 10−25 cm3), it can be
concluded that the 4F3/2 → 4I11/2 transition of NSASCF10
glass is suitable for a broadband optical amplification,
since it exhibits a relatively larger value of gain bandwidth
compared to reported glass systems such as NKZLSNd Fig. 4  Decay curves for the 4F3/2 level of different ­Nd3+ concentra-
tions in NSASCF glasses

Table 4  Emission peak positions (λp), effective bandwidths (∆λeff), radiative transition probabilities (AR), stimulated emission cross-section
(σemi), gain bandwidth parameter (σemi × ∆λeff) (× 10−25 cm3), experimental (βexp), and calculated branching ratios (βR) for the NSASCF10 glass
Transition 4F3/2 →  (λp) (∆λeff) (AR) (σemi) (σemi × ∆λeff) (βexp) (βR)
(nm) (nm) (s−1) (× 10−20cm2) (× 10−25 ­cm3)
4
I9/2 896 47 2813 2.00 0.94 0.39
4
I11/2 1058 35 3500 2.30 0.13 0.49
4
I13/2 1332 57 701 6.58 1.16 0.10
AT(s−1) 7014 2.03 0.63
τR (µs) 143 0.23

13
117 
Page 6 of 9 R. Doddoji et al.

where I(t) is the emission intensity at time t. The life- 3.5 SRIM and microscope images
times (τexp) of 4F3/2 level for 0.1, 0.5, 1.0, and 1.5 mol% of
NSASCF glasses are found to be 334 μs, 268 μs, 175 μs, and The simulation of the ion range for samples 1 and 2 is shown
104 μs, respectively. The lifetime is decreasing with increase in Fig. 5. From SRIM calculations, the carbon ions (sample
of ­Nd2O3 concentration due to the concentration quenching 1) are stopped at around 4 μm and the ion distribution is
through the energy transfer among ­Nd3+ ions [24, 36]. The very thin (a width of ~ 0.5 μm at half maximum). For sample
observed difference between βexp and βR values (see Table 4) 2, where protons were used, the ion range peak is located
may be due to experimental error and inaccuracy in the JO at ~ 6.3 μm. Protons penetrate deeper into the substrate, and
calculations; similar discrepancies have been reported in as the minimum available energy that could be used was
previously studied ­Nd3+-doped glasses [24, 36, 40]. 1 MeV, a 60° angle of incidence of the ion beam was used
For waveguide lasers, excitation of more active ions to obtain the chosen ion penetration. In addition, the use
­(Nd3+) is required in a small region of the glass matrix; of a higher angle than normal incidence generates a wider
thus, higher doping concentrations are preferable. There- ion distribution (the width was ~ 1.5 μm at half maximum).
fore, we selected NSASCF10 (sample 1) and NSASCF15 Images of the implanted zone are depicted in Fig. 6, from
(sample 2) glasses for the fabrication of planar waveguides which the measured damage depth for sample 1 is ~ 3 μm
in the present work. and for sample 2 is ~ 7 μm, which are in agreement with the
SRIM simulations.

3.6 Refractive index profiles

Sample 1 confines one propagation mode, and with help of the


radiation modes, the refractive index profile was approximated
and it is shown in Fig. 7a. This exhibits a refractive index
increase of ~ 0.012% in the guiding region and an optical bar-
rier located at ~ 3 μm with a decrease of ~ 0.1% with respect to
the substrate. Sample 2 propagates two modes, giving a refrac-
tive index profile with a waveguide width of ~ 7 μm, a slight
refractive index increase of ~ 0.006% in the guiding region, and
a barrier peak positioned ~ 0.25% with respect to the substrate
index value (see Fig. 7b). The refractive index increase in the
waveguide region has been reported previously and has been
related to ionization effects and stress produced by nuclear
collisions [41, 42]. Heavy ion implantation (e.g., carbon ions)
generates waveguides of a few microns with a higher energy
than that needed for light ions (e.g., protons). In addition,
Fig. 5  Ion range simulated with SRIM for the two glasses low heavy ion fluences (­ 1014 ions/cm2) normally ensure the

Fig. 6  Microscope images of the implanted zone: a sample 1 and b sample 2

13
Spectroscopic and waveguide properties of ­Nd3+‑doped oxyfluorosilicate glasses Page 7 of 9  117

(a) (b)

Fig. 7  Refractive index profiles for the planar waveguides: a sample 1 and b sample 2

formation of an optical barrier, which confines the light with- (which considers mode size mismatch and misalignment),
out deleterious effects in the waveguiding and spectroscopic Tl and Ts are the Fresnel reflection coefficients for lens and
properties [43–45]. In contrast, higher light ion fluences are sample, respectively, and L is the guide length in cm. Even
necessary to produce adequate light confinement; these are though the method does not determine each type of loss
generally in the order of ­1016 ions/cm2. According to the pre- separately (absorption, scattering, and tunnelling), the setup
vious reports [15, 16, 43, 44], the spectroscopic properties is simple and it gives a good approximation of the total prop-
are maintained in the waveguide region compared to the bulk agation losses per unit length.
substrate with the ion fluences used in this work; hence, similar Figure 8 shows the mode distribution at the waveguide
characteristics are expected in the waveguides. output of sample 2, where light confinement in the vertical
direction can be observed, defining a planar waveguide.
3.7 Waveguide propagation losses and mode Propagation losses for sample 1 were calculated in the
distribution range of 1.1–2.8 dB cm−1, and for sample 2 from 4.9 to
7 dB cm−1, measured in different regions along the wave-
The propagation losses in the waveguides were calculated guide output. From these values, it can be observed that
using the transmitted light method [15, 46] in which light is transmission properties depend on the implanted ion, that
measured at the input and output of the waveguide, and the is, the carbon implanted waveguide shows higher trans-
attenuation coefficient α is calculated using the following mission than the proton implanted waveguide [44]. In
expression: addition, losses caused by absorption could be reduced
( ) by applying an annealing treatment, which is commonly
10 P0 cTls Ts2 used in ion-implanted waveguides [15, 16].
𝛼= log dB/cm (5)
L P1

where P0 and P1 are the incident and output optical power,


respectively, c is the lens-waveguide coupling coefficient

Fig. 8  Waveguide mode inten-


sity distribution of sample 2

13
117 
Page 8 of 9 R. Doddoji et al.

4 Conclusions 11. G.Y. Chen, F. Piantedosi, D. Otten, Y.Q. Kang, W.Q. Zhang, X.
Zhou, T.M. Monro, D.G. Lancaster, Sci. Rep. 8, 10377 (2018).
https​://doi.org/10.1038/s4159​8-018-28631​-3
Nd3+-doped NSASCF glasses were prepared by the melt 12. C.S.D. Viswanath, K.V. Krishnaiah, C.K. Jayasankar, Opt.
quenching technique and investigated through absorption, Mater. 83, 348 (2018)
near-infrared emission, decay measurements, and wave- 13. V.R. Rao, C.K. Jayasankar, Opt. Mater. 91, 7 (2019)
14. M. Reben, E.S. Yousef, M. Piasecki, A.A. Albassam, A.M.
guiding properties. The amorphous nature of the glasses El-Naggar, G. Lakshminarayana, I.V. Kityk, I. Grelowska, J.
was confirmed by SEM analysis. EDS elemental mapping Mater. Sci. Mater. Electron 28, 8969–8975 (2017). https​://doi.
clearly depicts the incorporation of each concentration org/10.1007/s1085​4-017-6627-x
of ­Nd3+ ions into the glass host. Judd–Ofelt parameters 15. P.D. Townsend, P.J. Chandler, L. Zhang, Optical Effects of Ion
Implantation (CUP, Cambridge, 1994)
show the general trend Ω 6 > Ω4 > Ω2. The effect of con- 16. F. Chen, X.-L. Wang, K.-M. Wang, Opt. Mater. 29, 1523 (2007)
centration on the luminescence properties of ­Nd3+ ions 17. S. Stanek, P. Nekvindova, B. Svecova, S. Vytykacova, M. Mika,
has been discussed. Furthermore, the high stimulated J. Oswald, A. Mackova, P. Malinsky, J. Spirkova, Nucl. Instrum.
emission cross-section (6.58 × 10−20 cm2) and large gain Methods Phys. Res. B 371, 350 (2016)
18. X.-L. Shen, Q.-F. Zhu, R.-L. Zheng, P. Lv, H.-T. Guo, C.-X. Liu,
bandwidth (2.3 × 10 −25 cm3) for the 4F3/2 → 4I11/2 transi- Results Phys. 8, 352 (2018)
tion at 1058 nm are promising characteristics for amplifier 19. J.-Y. Chen, J.-Y. Lv, Z.-Y. Wang, S.-B. Lin, H.-T. Guo, C.-X.
and laser applications. In addition, the optical waveguides Liu, Results Phys. 12, 357 (2019)
exhibited good optical confinement and relatively low 20. C.X. Liu, B. Peng, W. Wei, S. Cheng, W.N. Li, H.T. Guo, Nucl.
Instrum. Methods Phys. Res. B 295, 85 (2013)
propagation losses. The combined properties of the glass 21. C.X. Liu, J. Xu, X.L. Xu, S. Wu, W. Wei, H.T. Guo, Opt. Eng.
host together with those of the optical waveguides show 53(3), 037101 (2014)
potential for the development of compact integrated opti- 22. 〈http://www.srim.org〉. Accessed 15 July 2018
cal devices. Further work will be carried out to test laser 23. P.J. Chandler, F.L. Lama, Opt. Acta 33, 127 (1986)
24. D. Ramachari, L.R. Moorthy, C.K. Jayasankar, Infra Phys. Tech-
emission performance in these glass waveguides. nol. 67, 555 (2014)
25. B.J. Chen, L.F. Shen, E.Y.B. Pun, H. Lin, Opt. Mater. Express
Acknowledgements  We are grateful to the optical workshop at CIO for 5(1), 113 (2015)
cutting and polishing the samples, as well as to K. López and F. Jaimes 26. S.S. Wang, Y. Zhou, Y.L. Lam, C.H. Kam, Y.C. Chan, X. Yao,
for supervising the implants. We would also like to thank Yogi Vemana Mater. Res. Innov. 1, 92 (1997)
University, Kadapa, for SEM and EDS measurements. 27. B.R. Judd, Phys. Rev. 127, 750 (1962)
28. G.S. Ofelt, J. Chem. Phys. 37, 511 (1962)
Compliance with ethical standards  29. D. Umamaheswari, B.C. Jamalaiah, T. Sasikala, G.V.L. Reddy,
L.R. Moorthy, J. Rare Earths 30-5, 413 (2012)
30. J.H. Campbell, T.I. Suratwala, J. Non Cryst. Solids 263&264,
Conflict of interest  The authors declare that they have no conflict of
318 (2000)
interest.
31. E.V. Uhlmann, M.C. Weinberg, N.J. Kredle, L.L. Burgner, R.
Zanohi, K.H. Church, J. Non Cryst. Solids 178, 15 (1994)
32. M. Ajroud, M. Haouari, H.B. Ouada, H. Maaref, A. Brenier, C.
Garapon, J. Phys. Condens. Matter 12, 3181 (2000)
33. Y. Tain, J. Zhang, X. Jing, S. Xu, Spectrochim. Acta Part A 98,
References 355 (2012)
34. T. Suzuki, H. Nasu, M. Hughes, S. Mizuno, K. Hasegawa, H.
1. S.L. Li, P.G. Han, M. Shi, Y.C. Yao, B. Hu, M.W. Wang, X.N. Ito, Y. Ohishi, J. Non Cryst. Solids 356, 2344 (2010)
Zhu, Opt. Express 19(24), 23958 (2011) 35. Y. Chen, Y. Huang, M. Huang, R. Chen, Z. Luo, J. Am. Ceram.
2. M.A.S. de Oliveira, C.B. de Araújo, Y. Messaddeq, Opt. Express Soc. 88, 19 (2005)
19(6), 5620 (2011) 36. P. Manasa, D. Ramachari, J. Kaewkhao, P. Meejitpaisan, E.
3. X. Li, G. Aka, L.H. Zheng, J. Xu, Q.H. Yang, Opt. Mater. Kaewnuam, A.S. Joshi, C.K. Jayasankar, J. Lum. 188, 558
Express 4(3), 458 (2014) (2017)
4. Y. Tian, J. Zhang, X. Jing, S. Xu, Spectrochim. Acta A Mol. 37. K.U. Kumar, V.A. Prathyusha, P. Babu, C.K. Jayasankar, A.S.
Biomol. Spectrosc. 98, 355 (2012) Joshi, A. Speghini, M. Bettinelli, Spectrochim. Acta A 67, 702
5. Y. Tian, R.R. Xu, L.L. Hu, J.J. Zhang, J. Appl. Phys. 111(7), (2007)
073503 (2012) 38. M. Venkateswarlu, S.K. Mahamuda, K. Swapna, M.V.V.K.S.
6. T. Suzuki, H. Kawai, H. Nasu, S. Mizuno, H. Ito, K. Hasegawa, Prasad, A.S. Rao, A.M. Babu, S. Shakya, G.V. Prakash, Opt.
Y. Ohishi, J. Opt. Soc. Am. B 28(8), 2001 (2011) Mater. 39, 8 (2015)
7. A.R. Molla, A. Tarafder, S. Mukherjee, B. Karmakar, Opt. 39. W.J. Miniscalco, J. Lightwave Technol. 9, 234 (1991)
Mater. Express 4(4), 843 (2014) 40. R. Balakrishnaiah, R. Vijaya, P. Babu, C.K. Jayasankar, M.L.P.
8. U. Skrzypczak, C. Pfau, C. Bohley, G. Seifert, S. Schweizer, J. Reddy, J. Non Cryst. Solids 353, 1397 (2007)
Phys. Chem. C 117(20), 10630 (2013) 41. L. Zhang, P.J. Chandler, P.D. Townsend, S.J. Field, D.C. Hanna,
9. A. Miguel, J. Azkargorta, R. Morea, I. Iparraguirre, J. Gonzalo, D.P. Shepherd, A.C. Tropper, J. Appl. Phys. 69, 3440 (1991)
J. Fernandez, R. Balda, Opt. Express 21(8), 9298 (2013) 42. G.V. Vázquez, Advances in quality of optical waveguides in
10. H. Lin, E.Y.B. Pun, X.R. Liu, J. Non Cryst. Solids 283, 27 Nd:YAG and ­LiNbO3, PhD Thesis, University of Sussex, 2000
(2001) 43. G.V. Vázquez, D. Ramírez, H. Márquez, E. Flores-Romero, J.
Rickards, R. Trejo-Luna, Appl. Opt. 51, 5573 (2012)

13
Spectroscopic and waveguide properties of ­Nd3+‑doped oxyfluorosilicate glasses Page 9 of 9  117

44. E. Flores-Romero, G.V. Vázquez, H. Márquez, R. Rangel-Rojo, Publisher’s Note Springer Nature remains neutral with regard to
J. Rickards, R. Trejo-Luna, Opt. Express 15, 8513 (2007) jurisdictional claims in published maps and institutional affiliations.
45. C.-X. Liu, J. Xu, L.-L. Fu et al., Opt. Eng. 54(6), 067106 (2015)
46. S.I. Najafi, Introduction to Glass Integrated Optics (Artech House,
Norwood, 1992)

13

S-ar putea să vă placă și