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

Control of the radiative properties via photon-plasmon interaction in Er3+-Tm3+-codoped tellurite glasses in the near infrared region

Open Access Open Access

Abstract

The novelty of this paper is that it reports on the tuning of the spectral properties of Er3+-Tm3+ ions in tellurite glasses in the near-infrared region through the incorporation of silver or gold nanoparticles. These noble metal nanoparticles can improve the emission intensity and expand the bandwidth of the luminescence spectrum centered at 1535 nm, covering practically all the optical telecommunication bands (S, C + L and U), and extended up to 2010 nm wavelength under excitation by a 976 nm laser diode. Both effects are obtained by the combined emission of Er3+ and Tm3+ ions due to efficient energy transfer processes promoted by the presence of silver or gold nanoparticles for the (Er3+)4I11/2→(Tm3+)3H5, (Er3+)4I13/2→(Tm3+)3H4 and (Er3+)4I13/2→(Tm3+)3F4 transitions. The interactions between the electronic transitions of Er3+ and Tm3+ ions that provide a tunable emission are associated with the dynamic coupling mechanism described by the variations generated by the Hamiltonian HDC in either the oscillator strength or the local crystal field, i.e. the line shape changes in the near-infrared emission band. The Hamiltonian is expressed as eigenmodes associated with the density of the conduction electron generated by the different nanoparticles through its collective free oscillations at each resonance frequency of the nanoparticle and their geometric dependence. A complete description of photon-plasmon interactions of noble metal nanoparticles with the Er3+ and Tm3+ ions is provided.

© 2014 Optical Society of America

1. Introduction

There has been an intense scientific and technological interest in the development of new materials as near-infrared (NIR) super-broadband and intense luminescence sources for the advancement of optical amplifiers technology, tunable lasers and amplified signals transmission [13]. Rare-earth (RE) ion -doped glasses are an alternative solution since the demand for optical devices that operate in the NIR region has been increasing.

Bulk glasses with high solubilities of RE ions represent a large optical confinement (medium gain [2,4,5]), which offers an attractive way of producing transparent, small and high-emission quantum yield devices. In contrast, the glasses fabrication with low contaminants content is still limited by the traditional technique of melting the powders precursor in crucible, despite the recent progresses observed in this area [6]. Specific properties of the host matrix, such as high refractive index, large optical transmission (visible and NIR regions) and low phonon energies are required for the achievement of high emission yield. In this sense, tellurite glasses (whose major component is TeO2) have appeared as optimal candidates among the oxide glasses because of their wide transparency range (from 400 to 5500 nm), high refractive index (~1.8-2.0), relatively low phonon energy (~800 cm−1) in comparison to silicate glasses [79] and good chemical and mechanical resistance [10,11]. Nevertheless, there exists a technological limitation to the use of RE ions for optical emission and amplification, since each RE ion covers the O-, E-, S-, C-, L- and U- bands of the optical communications separately (in the 1000 - 1680 nm region) [1,12]. The emission bands of the trivalent RE ions are difficult to surpass since the f-f transitions of the 4f orbitals are confined in the inner-shell and their nature is usually mediated by the electric dipole transitions which are sensitive to local environmental [13,14]. Several approaches are currently under development for a full use of the optical bandwidth through glassy systems, such as (i) doping with metal ions, e.g., bismuth [1518], nickel [1921], bismuth-nickel codoped [22], and, chromium [23], (ii) generation of supercontinuum light in highly non-linear optical fibers [2426] and (iii) co-doping with RE ions [2729]. Therefore, the expansion in the gain bandwidth in glassy materials for the manufacture of more efficient optical devices in the transmission network is a key factor for the present and future development of optical communications.

Noble metal nanoparticles (NPs) embedded in RE-doped tellurite glasses can either improve or quench the RE spectroscopic properties for optical applications in telecommunication bands [3034] and light emission by upconversion processes [3537]. The full understanding of the interactions involved in NPs and RE ions is still under debate. Some researchers have ascribed such an emission enhancement to the energy transfer (ET) between NP:RE, while others have attributed it to an improvement in the local field near the NPs (which exponentially decays from the NP surface). In fact, both interaction effects are produced from a coherent collective oscillation of the surface electrons of the metal NP (nonpropagating excitations), also known as localized surface plamon resonance (LSPR) [38,39], which enables the confinement of the electromagnetic energy (Ep=(ne2/(mε0))0.5 [39]) into volumes smaller than the diffraction limit (λ0/n)3, where ε0is the permittivity of the free space, n is the electron density, e and m are the charge and mass of the electron, respectively, n is the refractive index of the surrounding medium and λ0 is the incident excitation wavelength. Moreover the polarization of the NP can be either longitudinal (electric dipole parallel to the NP:RE axis) or transverse (electric dipole perpendicular to the NP:RE axis), depending on the NP shape, e.g. non-spherical NP, nanorods and others [4043]. Such electric dipole coupling is formed when both NP:RE have the same potential, but opposite polarity [38]. The controlled production of anisotropic NPs embedded in glassy materials [44], like tellurite glasses [3037,43] and the comprehension of the basic concepts involved constitute a current challenge that may bring us new technological applications.

The present research is focused on the emission enhancement and widening of broadbands in the NIR region in Er3+-Tm3+-doped tellurite glass by means of gold or silver NPs embedded in the glass for gain amplification. Such an emission bandwidth is obtained under excitation at 976 nm laser diode through an integrated emission between the Er3+ and Tm3+ ion levels via efficient ET processes: (Er3+)4I11/2→(Tm3+)3H5, (Er3+)4I13/2→(Tm3+)3H4 and (Er3+)4I13/2→(Tm3+)3F4, which cover all the S-, C + L- and U-bands of the optical telecommunications. These ETs are mediated by either silver or gold NPs (SNPs or GNPs). The NIR emission in the 1450-2010 nm region is ascribed to the following electronic transitions of Er3+ ions: 4I13/24I15/2 (1530 nm), 2H11/24I9/2 (1680 nm) and 4S3/24I9/2 (1720 nm), and the electronic transitions of Tm3+ ions: 3H43F4 (1490 nm) and 3F43H6 (1810 nm). Meanwhile, the NPs produce the following effects: (i) increment in the ET quantum yield, resulting in an emission improvement, and (ii) variations in the oscillator strength (local crystal field), resulting in a blue/red band shift. The latter may change the line shape and broaden the NIR emission band. Furthermore, the most outstanding characteristic of glassy materials and more particularly tellurite glasses has arisen from their potential applications as optical fiber amplifiers or fiber laser covering the entire telecommunication window.

2. Experimental section

Glasses of 74TeO2-5ZnO-15Na2O-5GeO2-1Er2O3:xTm2O3 (mol%) nominal composition with x = 0.05 or 0.06 mol% and containing 0.25 mol% of Ag or Au added in the form of chloride (AgCl or AuCl3) were prepared by the traditional melt-casting technique. More details about the preparation are provided in a previous paper [29]. For clarity, the studied glass samples are labeled as TxMy, where M represents the noble metal (S or G for silver and gold, respectively), x = 5 or 6 for the concentration of Tm2O3 (0.05 or 0.06 mol%, respectively), and y is the annealing time at 300 °C (y = 2, 4, 6, 8, 10 and 12 hours) – for the samples without NPs, i.e. 5 hours. All the glass samples were prepared from high purity starting materials (3N and above). Silver or gold NPs were generated in the host matrix by thermal mobility according to the annealing time after the decomposition of the metal chloride during the glass melting [3133,43,45]. The final samples were cut into 10 × 30 × 1 mm3 square pieces and polished until obtaining adequate transparency for the optical characterizations.

Dark-field microscopic images were obtained by an optical microscope (Olympus BX53 with a Darkfield dry condenser 10 × NA 0.8) coupled to a digital monochrome camera (Olympus XM10) and using a 75 W Xe lamp as a light source. A transmission dark-field objective (Olympus UPlanSApo 60x/1.2W NA 0.9 - NPLan) was used to focus the image of NPs embedded in the glasses.

The thermal properties of the glass samples were examined by differential scanning calorimetry (Netzsch DSC 404F3 Pegasus) in sealed Al pans at 10 °C/min heating rate. The density of the samples was measured by an MD-300S (Alfa Mirage) electronic densimeter and the Archimedes method using distilled water as the immersion liquid. Multiple density measurements were performed for each sample composition so as to obtain accurate values ( ± 5 mg/cm3 precision).

The transmission spectra were recorded from 350 to 2000 nm on a Varian Cary 500Scan UV-VI-NIR double beam spectrophotometer of ± 0.3 nm resolution. Steady-state luminescence spectra were obtained from a Nanolog spectrofluorimeter from Horiba Jobin Yvon equipped with a liquid-nitrogen-cooled Symphony InGaAs near-infrared detector. A pig-tailed diode laser at 976 nm (power 40 mW) was used as an external excitation source. The spectral slit width was 10 nm for the emission signal and the data acquired were corrected by instrumental factors. The samples were excited by a Pico Quant pulsed diode laser at 972 nm (model LDH-P-C) with pulse train of 60 ps and dead time of 20 ms for the measurements of the 4I13/2 excited state lifetime. The time resolved emission signal centered at 1535 ± 4 nm was measured by a Hamamatsu NIR-PMT module detector coupled to the Nanolog system by employing the method of Time Correlated Single-Photon Counting (TCSPC). Decay analysis software (DAS6) was used for obtaining the lifetime values by fitting a single exponential function. Measurements of both steady-state emission and lifetime were taken at room temperature. The laser beam was always collimated and focused at the center of the glass surface with a lens of 50 mm focal length.

3. Results and discussion

The glass transition temperatures Tg, determined for T5 and T6 samples are Tg = 311 ± 2°C and 309 ± 2°C, respectively. The annealing temperature used to generate and grow the NPs in both glasses was below their glass transition temperature Tg, at 300 °C. Such an experimental procedure was chosen so as to prevent a possible interaction between the glass network and the NPs mobile precursors [46], and a slow diffusion of the metallic Ag0 or Au0 particles [33]. Besides, the glass transition temperatures of the samples containing NPs were also determined: Tg = 303 ± 2°C and Tg = 301 ± 2°C for T5Sy and T6Sy samples, respectively, and Tg = 305 ± 2°C and Tg = 304 ± 2°C for T5Gy and T6Gy samples, respectively. The measured density of the samples, the calculated number density N for the Er3+, Tm3+, Ag0 and Au0 species, and the average distance ra-b between them [33,43] are summarized in Table 1. The values of glass densities, NEr3+, NTm3+, NAg0, NAu0, rAg0-Ag0, rAu0-Au0, rEr3+-Ag0 and rEr3+-Au0 are almost invariant for all the samples studied. The average distance rEr3+-Tm3+ in both glasses without NP is the same. Distance rEr3+-Tm3+ increases similarly upon the addition of metallic NPs (for both T5My and T6My samples). Likewise, from the values reported in Table 1, we can deduce the following trends:

Tables Icon

Table 1. Measured and calculated physical properties of TxMy glass samples: glass density, number (ionic or atomic) density N and average distance between rare-earth-ions and metallic-atoms (M0).

rEr3+Ag0<rEr3+Au0forallthesamples
rTm3+Au0<rTm3+Ag0forallthesamples

Almost the same rEr3+-M0 distance for the T5Gy and T6Gy, and T5Sy and T6Sy samples and a clear increase in the rTm3+-Ag0 distance from the T5Sy and T6Sy samples are observed. The Ag0 and Au0 atoms are closer to the Er3+ ions than to the Tm3+ ions, and the Tm3+ ions are closer to the Au0 atoms than to the Ag0 atoms. Besides, an increased rTm3+-M0 distance with the increases of the NTm3+ was observed. We have assumed that each Ag0 or Au0 atom has become a nucleation center and formed one NP during the annealing process (the growth is governed by the glass viscosity and the Ag0 or Au0 atoms mobility [45]) of specific size and shape. To maintain a thermodynamic and electronic charge equilibrium we can assume that the separation distance between RE and NP remains similar to the distance between RE ions and Ag0/Au0 atoms.

In summary, we have observed that: (i) the introduction of the metallic NPs increases the separation distance between the RE ions; (ii) the separation distances between Er3+ ions and SNP are lower than those of Er3+ ions and GNP, independently of the Tm3+ ions density NTm3+, suggesting a larger interaction (coupling) of SNP:Er3+ dipoles than GNP:Er3+, and in contrast; (iii) the separation distances between Tm3+ ions and SNP are larger than those of Tm3+ ions and GNP, suggesting a larger interaction of GNP:Tm3+ dipoles in comparison to SNP:Tm3+. Dark-field illumination is a microscopy technique largely used by biologists to observe unstained samples by making them appear brightly lit against a dark background [47]. The light scattered (transmitted) from the NP is directed to a camera image plane through an aperture of 0.1 mm diameter so as to reduce scattered light from other regions of the glass. The dark-field microscopic images obtained from T5S12 and T5G12 samples are shown in Fig. 1. From this, we can be observed that the size/shape and distribution those NPs are non-homogeneous inside the glasses, in agree with the literature [3037]. Although the GNPs or SNPs embedded in the glass do not show any fluorescence, the surface plasmon scattering they produce can be detected as a white light. The absence of fluorescence can be explained by the absorption of both Er3+ and Tm3+ ions present in the glass, which results in the colors emission turn-off of those NPs.

 figure: Fig. 1

Fig. 1 Optical dark-field microscopic images of the NPs embedded in T5S12 (a) and T5G12 (b) samples.

Download Full Size | PDF

The RE ions absorption bands can indeed overlap the NPs resonance wavelength, as verified in the glasses absorption spectra shown in Fig. 2. Nonetheless, from these dark-field images the different shapes of the NPs (spherical, elliptic and more complex) can be distinguished. Such an unorganized growth was predicted in [45], which demonstrates the optical properties of the NPs embedded in the glass can be explored.

 figure: Fig. 2

Fig. 2 Absorption spectra of (a) T5Sy, (b) T6Sy, (c) T5Gy, and (d) T6Gy glass samples. The inset shows a zoom of the absorption bands of transitions (a) 3H63F4, (b) 3H63H5, (c) 3H63H4 and (d) 3H63F2-3 in Tm3+ ions. All spectra show a high absorption of the Er3+ ions without NPs in comparison with the other samples, since the NPs increase the absorption these glasses.

Download Full Size | PDF

The room temperature absorption spectra of the TxMy samples are shown in Fig. 2. Eight intense transitions from the ground state to excited states 4I15/2→(4I13/2, 4I11/2, 4I9/2, 4F9/2, 4S3/2, 2H11/2, 4F7/2 and 4F3/2) are attributed to the presence of Er3+ ions and correspond to the bands centered at 1538, 978, 806, 649, 544, 520, 488 and 444 nm, respectively. Five subtle transitions from the ground state to excited states 3H6→(3F4, 3H5, 3H4, 3F3 and 3F2) are also observed for the Tm3+ ions and correspond to the bands centered at 1750, 1190, 785, 660 and 651 nm, respectively. Due to the small concentration of silver and gold (small atomic density NAg0 or NAu0) and the overlap of the LSPR band with the Er3+ or Tm3+ ions absorption bands, no plasmon band is observed in the absorption spectra recorded in the TxMy glasses. Nevertheless, the activity of the plasmon band associated with the presence of GNPs or SNPs in tellurite glasses was demonstrated in our previous investigations into the growth and geometry dependence on the optical properties of LSPR modes [31,32,45].

The luminescence spectra of the glasses in the 1400-2120 nm range under 976 nm laser diode excitation (with a diode power of 40 mW) are shown in Fig. 3. Note that the measurement reproducibility was carefully verified for all samples studied.

 figure: Fig. 3

Fig. 3 Near infrared luminescence spectra under excitation at 976 nm of (a) T5Sy, (b) T6Sy, (c) T5Gy, and (d) T6Gy glass samples showing the 4I13/24I15/2 and 3F43H6 radiative transitions of the Er3+ and Tm3+ ions, respectively, and their splitting due to the Stark effect.

Download Full Size | PDF

The oscillator strength (ffi) of the trivalent RE ion is a fundamental physical quantity in analytical spectroscopy which determines the sensitivity of the 4f-4f transitions due to the chemical nature and symmetry of the ligand environment. This phenomenon is referred to as hypersensitivity transition and has been correlated with the coordination numbers and symmetries of the RE ions in the host matrix. On the other hand, LSPR creates electromagnetic resonance fields that can improve the local field [38]. Such fields are formed through oscillating dipoles (from the NPs) around the RE ion, hence an additional oscillating electric field produced near the RE ion, which may induce strong changes in 4f–4f transitions [31] (e.g. a blue/red band shift or an enhanced/quenched luminescence). The induced oscillating dipoles depend on the size, shape and isotropic dipolar polarizabilities of the NPs environment. Therefore, evident variations occur in the luminescence spectra due to the NPs embedded in the glass, Fig. 3. In the first approximation, we can use the dynamic coupling mechanism [4850] and the interaction energy between the NP:RE is given by: HDC=ei,jμjnjriRj|riRj|3, whereμj(=αjEinc) is the electric dipole moment induced in the NP by an incident electric field Einc, |riRj|is the separation distance between NP:RE, nj is the density of the conduction electrons inside a plasma characterized by carriers with charge N’e.

According to our experimental conditions, we must consider that: (i) as the growth of the NPs is governed by the thermal treatment, NPs can have different geometrical ratios (e.g. ellipsoidal NPs), therefore the polarizability is [51] αj=4π3a1a2a3εNPε(ω)ε(ω)Li(εNPε(ω)), where principal axes a1, a2 and a3 introduce geometrical depolarization factors Li (for spherical particles, L1 = L2 = L3 = 1/3, or spheroidal particles L1 = L2), ε(ω) and εNP are, respectively, the dielectric and NP permittivity as a function of the excitation frequency; (ii) we can assume a couple of particles (NP and RE ions) forming an electric dipole of −V/2 and + V/2 potentials, respectively [33]; (iii) the ET between Er3+:Tm3+ is mediated by the NPs [38,43]; (iv) the ET from the RE ion to NP is more probable than the ET from NP to the RE ion, since the excited state lifetime of NPs is extremely shorter (ns order) in comparison with the higher energy excited states of Er3+ and Tm3+ ions (μs or ms order) [29,52]; (v) the luminescence quenching can be described by the ET between NP:RE and re-absorption by the NPs which are in resonance with the emissions of RE ions [33,38,43]. Such considerations are summarized in Fig. 4. About the point (ii), in the case with multiple interactions at different distances between NPs:REs, the charge from the RE ion is the same, because the Eincemployed to excite these particles will be absorbed for the RE ion due to the cross-section absorption those is more than the NPs [Fig. 2], case contrary the light incident (λ0≠λspp – wavelength surface plasmon resonance) is scattered for the NPs, and this scattered light is absorbed for the RE ion which emits and this coupled/resonates with the NP.

 figure: Fig. 4

Fig. 4 Schematic representation of the Er3+-Tm3+:NP interactions in two different scenarios: (a) the RE ions do not interact with each other, but with the NP and; (b) the interaction occurs between NP:RE and RE1:RE2. Equipotential surface (electric multipole coupling) with electric potentials −V/2 and + V/2 between NP:REs,d=1,2|riRj|.

Download Full Size | PDF

Several interesting aspects about the oscillator strength obtained by dynamic coupling mechanisms between NP:RE can be discussed. We can write the transition probability from any excited state transition decaying to background (2→1) as A21=8π2e2λ2mcf21, where c is the light velocity. The intensity emitted from a glass can be expressed by Iem=hυN2A21, where N2 is the population in the excited level, h is the Planck constant and υ is the line frequency, and we have the following relation Iemf21. In this frame, the NPs contribute only to f21 when Eincis present and the states of the REs have opposite parity [31,38]. Figure 3 shows the magnitude of the HDC effects on the crystalline potential as changes in the spectral lines, their widths, and shapes. In glasses, since different sites can be occupied by the RE ion, the Stark splitting is not well determined. However, here the manufactured process was strictly the same for all the samples, therefore the Stark splitting must be invariant in our samples. In this manner, the presence of NPs in the host matrix can modify the bandwidth of the Stark energy levels, which results in a blue/red shift and/or a broadening/increase in the RE ions emission as well as the emission intensity.

A subtle increase in NTm3+ induces a decrease in the emission intensity of the Er3+ and Tm3+ ions due to the ET from Er3+ to Tm3+ ions (donor-acceptor) and a cross-relaxation process in which the quenching reduces the luminescence quantum efficiency of the Er3+-Tm3+ ions [Fig. 4]. A similar behavior was observed by V.A.G. Rivera [29]. Besides, such ET process between the Er3+ to Tm3+ ions still remains in the TxMy samples since the separation distance between the RE ions does not change with decreasing Tm3+ ions concentration, Table 1.

On the one hand, in comparison to the T5 glass sample, the T5S08 sample shows 282% and 238% improvements in the emission intensity near 1535 and 1795 nm for the Er3+:(4I13/24I15/2) and Tm3+:(3F43H6) transitions, respectively. On the other hand, in comparison with the T6 sample, the T6S10 sample shows 233% and 248% improvements in the emission intensity for the same Er3+ and Tm3+ ions transitions. Such values of NIR emission enhancement induced by NPs have been reported for the first time in tellurite glasses. In order to analyze the oscillator strength ffi, the values of the intensity peaks (Iλ-peak) were taken from the emission spectra [Fig. 3], with and without NPs at 1535 ± 6 and 1795 ± 8 nm wavelengths [Fig. 5(a)]. The ± 6 or 8 nm deviations are due to the blue/red shift produced by the longitudinal/transverse polarization of the NPs [38,43]. Figure 5(a) provides an overview of the amplitude of the luminescence enhancement (↑) or quenching (↓) observed in the glass samples. The emission peak intensity strongly depends on the annealing time, and NPs nature and shape, therefore: (i) the peak intensity ratio Iλ-1535/Iλ-1795 varies from the glass without NPs (Tx) to the samples containing NPs (TxMy), which confirms the HDC effects; (ii) the T5S08 sample exhibits a more intense emission than the other TxMy samples, and the T5G10 sample shows a more intense emission than the other TxGy samples; (iii) Iλ-1535 > Iλ-1795 for the TxSy samples (except the T6S06 and T6S10 samples), indicating that the SNP:Er3+ coupling is more intense than SNP:Tm3+, which is in agreement with the average separation distance for these ions (rEr3+-Ag0 < rEr3+-Au0); (iv) in contrast, for TxGy samples, Iλ-1535 < Iλ-1795 (except for the T6G10 sample), indicating that the coupling of GNP is more intense with the Tm3+ ions than with the Er3+ ions, which is also in agreement with the average separation distance for these ions (rTm3+-Au0 < rTm3+-Ag0); (v) although most of the TxSy samples show higher emission intensities, some of them also show a serious quenching in comparison to the TxGy samples, see doted/dashed lines in Fig. 5(a). We can conclude that the multidipole SNP:(Er3+-Tm3+) displays a strong coupling due to a short separation distance, which can improve luminescence, or quench, due to the ET between SNP:RE or re-absorption by the SNPs, which are in resonance with the emissions of RE ions [43].

 figure: Fig. 5

Fig. 5 (a) Intensity of the emission peaks at 1535 ± 6 nm (4I13/24I15/2 transition of Er3+ ion) and 1795 ± 8 nm (3F43H6 transition of Tm3+ ion) for the TxMy samples studied. (b) Bandwidth of the emission band centered at 1535 ± 6 nm for the TxMy samples studied. (c) Emission spectra of the T5S08 and T6S10 samples which exhibit the most intense emission, and T5G08, T5G10, T6G02 and T6G12 samples which exhibit the broadest band.

Download Full Size | PDF

In a previous study on the expansion of broadband emissions in NIR [29], we achieved a covering of the optical telecommunications S-, C + L- and U-bands by optical pump at 976 nm (for the Tm0.075 sample in [29]) through an integrated emission by means of appropriate Er3+-Tm3+ co-doping ions and an ET among levels (Er3+)4I11/2→(Tm3+)3H5, (Er3+)4I13/2→(Tm3+)3H4 and (Er3+)4I13/2→(Tm3+)3F4. Whereas the sample labeled Tm0.050 in [29] is similar to the T5 sample of this study, the emission bandwidth observed here is larger than that of the Tm0.050 sample (112 nm vs. 79 nm, respectively). This difference may be attributed to the host glass composition (with different sodium and zinc oxide concentrations) and the annealing time. Therefore, the T6 sample shows a 118 nm bandwidth. The bandwidths of T5 and T6 samples [Fig. 5(c)], are less important than that of the Tm0.075 sample (134 nm bandwidth [29]).

The TxGy samples exhibit larger bandwidths in comparison to the other samples, of which the most promising are T5G08, T6G02, T5G10 and T6G12 with expanded broadbands of 411, 455, 464 and 467 nm, respectively, Fig. 5(b). To the best of our knowledge, this is the first time such large values of FWHM have been reported for an NIR emission band in tellurite glasses. The FWHM values were determined at 1535 ± 6 nm, covering the second part of the S-band (1460 ± 10 to 1530 nm), the entire C + L-band (1530-1625 nm) and the U-band (1625-1675 nm), and extended beyond 285 nm (up 1995 ± 10 nm). This broadband emission results from the combined emission of the Er3+ and Tm3+ ions through efficient ET processes mediated by the metallic NPs via HDC [Fig. 4], according to the following transitions: (Er3+)4I11/2→(Tm3+)3H5, (Er3+)4I13/2→(Tm3+)3H4 and (Er3+)4I13/2→(Tm3+)3F4, as shown in Fig. 6. The broadband emission results from the energy splitting of the Stark levels of each Er3+ ion (with 7 ( = (2J + 1)/2)) and Tm3+ ion (with 8 ( = 2J + 1)), the multipole interactions between NP:RE, and also the multipole interactions between the host matrix ions [29,38]. The NP:RE interactions were described through HDC, i.e. the perturbation theory is applicable here. We can indeed add a “perturbation series”, i.e. the “k” terms, since we have different separation distances di,j, shape and size of the NPs [Fig. 4], which are responsible for shifting and splitting the luminescence spectra, as shown in Fig. 5(c), due to dynamic coupling mechanism. Therefore, the states have different crystal-field representations, and more than one type of absorption/emission dipole can be simultaneously excited or emitted.

 figure: Fig. 6

Fig. 6 Schematic energy level diagrams of (a) TxSy samples and (b) TxGy samples. Ground state absorption (GSA) of the Er3+ under 976 nm excitation resulting in a population inversion of the high (via ESA) and lower levels through nonradiative (NR) decays. Radiative ET among Er3+-Tm3+ ions (ETiE-iT), nonradiative energy transfer ETNR, and ET between the NP:RE (ETSNP or ETGNP). τNP is the finite plasmon lifetime (ns order) and the red curved lines represent the displacement of a small volume charge nj=qp,j(r)eiωpt to the quantum system.

Download Full Size | PDF

In this scenario, we can use a modal approach similar to that developed in recent quantum theories that describes the propagation of surface plasmon created from the LSPR [53,54]. We consider displacements varying in time as harmonic (damped) oscillators, i.e. nj=nj(r)eiωt. The optical properties of an NP are ruled by the collective excitation of its free electron cloud and create an electric dipole in the glass with polarizability αj. In this way, as described in [43], the ET between NP:RE could have either a positive or a negative impact for a specific electronic transition and depends on the efficiency of the coupling. Let us consider a set of orthonormal vector functions qp(r) (eigenfunctions) representing the amplitude of a harmonic displacement [54] that could be coupled with the RE electronic transitions. We can write nj=qj,p(r)eiωpt and finally HDC=ei,jμj,pqp,j(r)eiωptriRj|riRj|3.

The Hamiltonian HDC is expressed as a function of the set qj,p eigenmodes associated with the conduction electron density generated by different NPs (j) through its collective free oscillations at each resonance frequency ωp of the NP and their geometric dependence by means of αj,p. The quantum behavior of RE ions and plasmons (in LSPR mode) interacts through HDC through qj,p, resulting in a plasmon-photon coupling [Fig. 6], which can be verified as an enhancement/quenching of luminescence, a widening of the broadband emission and a change in its line shape, as observed in the luminescence spectra in Figs. 3 and 5.

Because of this process, the plasmon states on the RE ions are coupled, and the respective energies are modified, nevertheless, the occupation of a given plasmon mode on a given electronic level of the RE ion (Er3+ or Tm3+, Fig. 6) will depend of the efficiency of the ET between NP and RE. Furthermore, when the higher energy levels of Er3+ and Tm3+ ions are coupled with the NP (the collective free oscillations with each resonance frequencyωp), they can generate a white light, as initially observed in the NP in Fig. 1.

Figure 6 shows the schematic energy level diagrams in the vicinity of either an SNP [Fig. 6(a)] or GNP [Fig. 6(b)] of the Er3+-Tm3+ ions doped tellurite glass. The excitation at 976 nm stimulates the Er3+ ion from the ground level 4I15/2 to the 4I11/2 level and then to the 4F7/2 level through an excited state absorption process (ESA), in which the nonradiative decays populate the levels with lower energy of the Er3+ ion. Likewise, the ET process from Er3+ to Tm3+ ions can populate the other energy levels of Tm3+ ion and the levels of lower energy by nonradiative decays. Er3+:(4S3/2) and Tm3+:(3H4) excited states can also be populated by ET between two neighboring Er3+-Tm3+ ions in the present system. Besides, the plasmon-photon coupling has been drawn as a red curved line (ETNP) representing the displacement of a small volume charge nj,p=qj,p(r)eiωpt to the quantum system. Accordingly, the NIR emission in the 1450-2010 nm region can be unambiguously ascribed to the following transitions of the Er3+ ions: 4I13/24I15/2 (1530 nm), 2H11/24I9/2 (1680 nm) and 4S3/24I9/2 (1720 nm) and of the Tm3+ ions: 3H43F4 (1490 nm) and 3F43H6 (1810 nm), Fig. 6.

Another effect produced by the presence of NPs is the variation in the 4I13/2 lifetime, shown in Fig. 7, due to radiative/nonradiative ET after the incidence of the pump energy to excite the NPs. This means there occur an increment in the light absorption due to the presence of NPs [Fig. 2] and results in a lifetime increase in the T5My and T6Sy samples in comparison with the T5 and T6 samples, respectively. A decrease in the lifetime is observed for the T6Gy samples and more particularly for T5G08, T5G10, T6G02 and T6G12, which also exhibit an expanded broadband [↔, Fig. 7] and higher emission intensity [↑, Fig. 7]. In this manner, the increase or decrease of the lifetime is due to the presence of the NPs with a little or large damping (γ=1/τNP [33]), respectively, which favors or not the ET processes, Fig. 6. The T5G10 sample has shown improvements in both luminescence intensity and bandwidth broadening, which makes the sample an excellent candidate as an optical gain medium for potential applications in amplifying waveguides for plasmonic devices and other photonic nano-devices [4,55,56].

 figure: Fig. 7

Fig. 7 Lifetime of the Er3+:4I13/2 level for all samples studied. Dashed and dotted lines indicate, respectively, the lifetime of the sample without SNP and GNP for comparison. The T5G08, T5G10, T6G02 and T6G12 samples exhibit an expanded broadband (↔). The T5S08, T6S10 and T5G10 samples exhibit an improved luminescence (↑).

Download Full Size | PDF

Conclusions

Er3+-Tm3+-doped tellurite glasses are good host matrices with broadband NIR emissions of 112 and 118 nm under excitation at 976 nm for glasses co-doped with 1 mol.% of Er3+ and 0.05 and 0.06 mol% of Tm3+, respectively. For the first time the tuning in the spectral properties of Er3+-Tm3+ ions in tellurite glasses in the near-infrared region has been demonstrated through the incorporation of either SNP or GNP. A decrease in both emission intensity and lifetime of the 4I13/2 level of Er3+ ion was observed by increasing the Tm3+ concentration, due to energy transfers. These tellurite glasses have enabled the formation of SNP and GNP embedded in the host matrix of different shape and size as a function of the annealing time, as verified through dark-field microscopy. The presence of NPs induces the formation of electric multiple dipoles activated by the ET between the Er3+-Tm3+ ions and the NPs, forming LSPR. The separation distance between Er3+-Tm3+ ions increases with the presence of the NPs, the SNP is closer to the Er3+ ions and GNP is closer to Tm3+. In this framework, the dynamic coupling mechanism that describes the interaction energy between the NP:RE was used through the Hamiltonian HDC expressed as a function of the eigenmodes set. Such eigenmodes were associated with the conduction electron density generated by different NPs through its collective free oscillations at each resonance frequency of the NP and their geometric dependence. Therefore, changes in the Stark splitting, blue/red shift, and/or an increase in the bandwidth of the Er3+-Tm3+ NIR emissions can be described through the HDC. The T5S08 and T6S10 glass samples showed an increase in the intensity of the Er3+:(4I13/24I15/2) and Tm3+:(3F43H6) emissions in comparison with that of the other samples. Likewise, the T5G08, T5G10, T6G02 and T6G12 glass samples exhibited larger bandwidths in comparison to those of the other samples. Both emission enhancement and broadband widening are in agreement with the Er3+:4I13/2 lifetimes obtained under diode laser at 976 nm as the excitation source. To the best of our knowledge, such values of NIR emission from RE ions controlled via photon-plasmon interaction have been reported for the first time in tellurite glasses in this study. From a technological point of view, all the NIR luminescence features reported here, especially for the T5G10 sample and its NIR emission (473 nm bandwidth and enhancement of luminescence of approximately 44%) make these glasses promising candidates for broadband optical fiber amplifiers in the optical communication window and supercontinuum laser source. From a theoretical point of view, this study also brings fresh insight, since the Er3+-Tm3+-doped tellurite glasses with embedded metallic NPs have enabled us experimental study the photon-plasmon interactions, which have resulted in an improvement in the emission quantum yield of the RE ions.

Acknowledgements

This research was supported by Brazilian agencies CAPES, FAPESP – Processes: 2009/08978-4 and 2011/21293-0, and CNPq through the INOF/CEPOF (Instituto Nacional de Óptica e Fotônica and Centro de Pesquisa em Óptica e Fotônica – São Paulo - Brasil) and the Canadian Excellence Research Chair program (CERC) on Enabling Photonic Innovations for Information and Communication.

References and links

1. M. Yamane and Y. Asahara, Glasses for Photonics (Cambridge University, 2000).

2. K. Richardson, D. Krol, and K. Hirao, “Glasses for photonic applications,” Int. J. Appl. Glass Sci. 1(1), 74–86 (2010). [CrossRef]  

3. S. Tanabe, “Rare-earth-doped glasses for fiber amplifiers in broadband telecommunication,” C. R. Chim. 5(12), 815–824 (2002). [CrossRef]  

4. V. A. G. Rivera, Y. Ledemi, S. P. A. Osorio, F. A. Ferri, Y. Messaddeq, L. A. O. Nunes, and E. Marega Jr., “Optical gain medium for plasmonic devices,” Proc. SPIE 8621, 86211J (2013). [CrossRef]  

5. A. J. Kenyon, “Recent developments in rare-earth doped materials for optoelectronics,” Prog. Quantum Electron. 26(4-5), 225–284 (2002). [CrossRef]  

6. K. A. Varshneya, H. A. Richardson, M. Wightman, and L. D. Pye, Processing, Properties and Applications of Glass and Optical Materials (John Wiley, 2012).

7. J. S. Wang, E. M. Vogel, and E. Snitzer, “Tellurite glass: a new candidate for fiber devices,” Opt. Mater. 3(3), 187–203 (1994). [CrossRef]  

8. G. S. Maciel, C. B. de Araújo, Y. Messaddeq, and M. A. Aegerter, “Frequency upconversion in Er3+-doped fluoroindate glasses pumped at 1.48 μm,” Phys. Rev. B 55(10), 6335–6342 (1997).

9. A. S. Oliveira, M. T. de Araujo, A. S. Gouveia-Neto, J. A. Medeiros Neto, A. S. B. Sombra, and Y. Messaddeq, “Frequency up-conversion in Er3+/Yb3+-codoped chalcogenide glass,” Appl. Phys. Lett. 72(7), 753–755 (1998). [CrossRef]  

10. A. Jha, B. Richards, G. Jose, T. Teddy-Fernandez, P. Joshi, X. Jiang, and J. Lousteau, “Rare-earth ion doped TeO2 and GeO2 glasses as laser materials,” Prog. Mater. Sci. 57(8), 1426–1491 (2012). [CrossRef]  

11. R. A. H. El-Mallawany, Tellurite Glasses Handbook – Physical Properties and Data (CRC Press, 2001).

12. G. P. Agrawal, Fiber-Optic Communication Systems (John Wiley, 2002).

13. M. J. F. Digonnet, Rare-Earth-Doped Fiber Lasers and Amplifiers (TM Marcel Dekker, 2001).

14. S. Karaveli and R. Zia, “Spectral tuning by selective enhancement of electric and magnetic dipole emission,” Phys. Rev. Lett. 106(19), 193004 (2011). [CrossRef]   [PubMed]  

15. S. F. Zhou, N. Jiang, B. Zhu, H. C. Yang, S. Ye, G. Lakshminarayana, J. H. Hao, and J. R. Qiu, “Multifunctional Bismuth-doped nanoporous silica glass: From blue-green, orange, red, and white light sources to ultra-broadband infrared amplifiers,” Adv. Funct. Mater. 18(9), 1407–1413 (2008). [CrossRef]  

16. X. G. Meng, J. R. Qiu, M. Y. Peng, D. P. Chen, Q. Z. Zhao, X. W. Jiang, and C. S. Zhu, “Near infrared broadband emission of bismuth-doped aluminophosphate glass,” Opt. Express 13(5), 1628–1634 (2005). [CrossRef]   [PubMed]  

17. Y. Arai, T. Suzuki, Y. Ohishi, K. S. Morimoto, and S. Khonthon, “Ultrabroadband near-infrared emission from a colorless bismuth-doped glass,” Appl. Phys. Lett. 90(26), 261110 (2007). [CrossRef]  

18. M. Peng, X. Meng, J. Qiu, Q. Zhao, and C. Zhu, “GeO2: Bi, M (M=Ga, B) glasses with super-wide infrared luminescence,” Chem. Phys. Lett. 403(4-6), 410–414 (2005). [CrossRef]  

19. T. Suzuki and Y. Ohishi, “Broadband 1400 nm emission from Ni2+ in zinc—alumino—silicate glass,” Appl. Phys. Lett. 84(19), 3804–3806 (2004). [CrossRef]  

20. S. Zhou, N. Jiang, K. Miura, S. Tanabe, M. Shimizu, M. Sakakura, Y. Shimotsuma, M. Nishi, J. Qiu, and K. Hirao, “Simultaneous tailoring of phase evolution and dopant distribution in the glassy phase for controllable luminescence,” J. Am. Chem. Soc. 132(50), 17945–17952 (2010). [CrossRef]   [PubMed]  

21. S. Zhou, N. Jiang, B. Wu, J. Hao, and J. Qiu, “Ligand-driven wavelength-tunable and ultra-broadband infrared luminescence in single-ion-doped transparent hybrid materials,” Adv. Funct. Mater. 19(13), 2081–2088 (2009). [CrossRef]  

22. K. Zhang, S. Zhou, Y. Zhuang, R. Yang, and J. Qiu, “Bandwidth broadening of near-infrared emission through nanocrystallization in Bi/Ni co-doped glass,” Opt. Express 20(8), 8675–8680 (2012). [CrossRef]   [PubMed]  

23. S. Tanabe and X. Feng, “Temperature variation of near-infrared emission from Cr4+ in aluminate glass for broadband telecommunication,” Appl. Phys. Lett. 77(6), 818–820 (2000). [CrossRef]  

24. B. R. Washburn, S. A. Diddams, N. R. Newbury, J. W. Nicholson, M. F. Yan, and C. G. Jørgensen, “Phase-locked, erbium-fiber-laser-based frequency comb in the near infrared,” Opt. Lett. 29(3), 250–252 (2004). [CrossRef]   [PubMed]  

25. K. Saitoh and M. Koshiba, “Highly nonlinear dispersion-flattened photonic crystal fibers for supercontinuum generation in a telecommunication window,” Opt. Express 12(10), 2027–2032 (2004). [CrossRef]   [PubMed]  

26. M. El-Amraoui, G. Gadret, J. C. Jules, J. Fatome, C. Fortier, F. Desevedavy, I. Skripatchev, Y. Messaddeq, J. Troles, L. Brilland, W. Gao, T. Suzuki, Y. Ohishi, and F. Smektala, “Microstructured chalcogenide optical fibers from As2S3 glass: towards new IR broadband sources,” Opt. Express 18(25), 26655–26665 (2010). [CrossRef]  

27. B. Zhou, L. Tao, Y. H. Tsang, W. Jin, and E. Y. Pun, “Superbroadband near-infrared emission and energy transfer in Pr3+-Er3+ codoped fluorotellurite glasses,” Opt. Express 20(11), 12205–12211 (2012). [CrossRef]   [PubMed]  

28. S. X. Shen, A. Jha, L. H. Huang, and P. Joshi, “980-nm diode-pumped Tm3+/Yb3+-codoped tellurite fiber for S-band amplification,” Opt. Lett. 30(12), 1437–1439 (2005). [CrossRef]   [PubMed]  

29. V. A. G. Rivera, M. El-Amraoui, Y. Ledemi, Y. Messaddeq, and E. Marega Jr., “Expanding broadband emission in the near-IR via energy transfer between Er3+–Tm3+ co-doped tellurite-glasses,” J. Lumin. 145, 787–792 (2014). [CrossRef]  

30. M. Eichelbaum and K. Rademann, “Plasmonic enhancement or energy transfer? on the luminescence of gold-, silver-, and lanthanide-doped silicate glasses and its potential for light-emitting devices,” Adv. Funct. Mater. 19(13), 2045–2052 (2009). [CrossRef]  

31. V. A. G. Rivera, S. P. A. Osorio, Y. Ledemi, D. Manzani, Y. Messaddeq, L. A. O. Nunes, and E. Marega Jr., “Localized surface plasmon resonance interaction with Er3+-doped tellurite glass,” Opt. Express 18(24), 25321–25328 (2010). [CrossRef]   [PubMed]  

32. S. P. A. Osorio, V. A. G. Rivera, L. A. O. Nunes, E. Marega Jr, D. Manzani, and Y. Messaddeq, “Plasmonic coupling in Er3+:Au tellurite glass,” Plasmonics 7(1), 53–58 (2012). [CrossRef]  

33. V. A. G. Rivera, Y. Ledemi, S. P. A. Osorio, D. Manzani, Y. Messaddeq, L. A. O. Nunes, and E. Marega Jr., “Efficient plasmonic coupling between Er3+:(Ag/Au) in tellurite glasses,” J. Non-Crys. Sol. 358, 399–405 (2012).

34. V. A. G. Rivera, D. Manzani, Y. Messaddeq, L. A. O. Nunes, and E. Marega Jr., “Study of Er3+ fluorescence on tellurite glasses containing Ag nanoparticles,” J. Phys. Conf. Ser. 274, 012123 (2011). [CrossRef]  

35. M. Reza Dousti, M. R. Sahar, R. J. Amjad, S. K. Ghoshal, A. Khorramnazari, A. Dordizadeh Basirabad, and A. Samavati, “Enhanced frequency upconversion in Er3+-doped sodium lead tellurite glass containing silver nanoparticles,” Eur. Phys. J. D 66(9), 237 (2012). [CrossRef]  

36. R. P. Kassab Luciana, R. De Almeida, D. M. Da Silva, A. A. De Assumpção Thiago, and C. B. De Araújo, “Enhanced luminescence of Tb3+/Eu3+ doped tellurium oxide glass containing silver nanostructures,” J. Appl. Phys. 105, 103505 (2009).

37. . P. P. de Campos, L. R. P. Kassab, T. A. A. de Assumpção, D. S. da Silva, and C. B. de Araújo, “Infrared-to-visible upconversion emission in Er3+ doped TeO2-WO3-Bi2O3 glasses with silver nanoparticles,” J. Appl. Phys. 112(6), 063519 (2012).

38. V. A. G. Rivera, F. A. Ferri, and E. Marega, Jr., Plasmonics - Principles and Applications (InTech, 2012), Chap. 11.

39. K. M. Mayer and J. H. Hafner, “Localized surface plasmon resonance sensors,” Chem. Rev. 111(6), 3828–3857 (2011). [CrossRef]   [PubMed]  

40. S. Link and M. A. El-Sayed, “Spectral properties and relaxation dynamics of surface plasmon electronic oscillations in gold and silver nanodots and nanorods,” J. Phys. Chem. B 103(40), 8410–8426 (1999). [CrossRef]  

41. P. K. Jain, S. Eustis, and M. A. El-Sayed, “Plasmon coupling in nanorod assemblies: Optical absorption, discrete dipole approximation simulation, and exciton-coupling model,” J. Phys. Chem. B 110(37), 18243–18253 (2006). [CrossRef]   [PubMed]  

42. A. Abbas, L. Tian, J. J. Morrissey, E. D. Kharasch, and S. Singamaneni, “Hot spot-localized artificial antibodies for label-free plasmonic biosensing,” Adv. Funct. Mater. 23, 1789–1797 (2013).

43. V. A. G. Rivera, Y. Ledemi, S. P. A. Osorio, D. Manzani, F. A. Ferri, J. L. Ribeiro Sidney, L. A. O. Nunes, and E. Marega Jr, “Tunable plasmon resonance modes on gold nanoparticles in Er3+-doped germanium–tellurite glass,” J. Non-Crys. Solids 378, 126–134 (2013).

44. A. Simo, J. Polte, N. Pfänder, U. Vainio, F. Emmerling, and K. Rademann, “Formation mechanism of silver nanoparticles stabilized in glassy matrices,” J. Am. Chem. Soc. 134(45), 18824–18833 (2012). [CrossRef]   [PubMed]  

45. V. A. G. Rivera, S. P. A. Osorio, D. Manzani, Y. Messaddeq, L. A. O. Nunes, and E. Marega Jr., “Growth of silver nano-particle embedded in tellurite glass: Interaction between localized surface plasmon resonance and Er3+ ions,” Opt. Mater. 33(6), 888–892 (2011). [CrossRef]  

46. V. A. G. Rivera, Ion Exchange Technology I, (Springer, 2012), Chap. 14.

47. K. J. Savage, M. M. Hawkeye, R. Esteban, A. G. Borisov, J. Aizpurua, and J. J. Baumberg, “Revealing the quantum regime in tunnelling plasmonics,” Nature 491(7425), 574–577 (2012). [CrossRef]   [PubMed]  

48. C. K. Jørgensen and B. R. Judd, “Hypersensitive pseudoquadrupole transitions in lanthanides,” Mol. Phys. 8(3), 281–290 (1964). [CrossRef]  

49. M. F. Reid and F. S. Richardson, “Electric dipole intensity parameters for lanthanide 4f → 4f transitions,” J. Chem. Phys. 79(12), 5735–5742 (1983). [CrossRef]  

50. O. L. Malta and L. D. Carlos, “Intensities of 4f-4f transitions in glass materials,” Quim. Nova 26(6), 889–895 (2003). [CrossRef]  

51. S. A. Maier and H. A. Atwater, “Plasmonics: Localization and guiding of electromagnetic energy in metal/dielectric structures,” J. Appl. Phys. 98(1), 011101 (2005). [CrossRef]  

52. O. L. Malta and M. A. Couto dos Santos, “Theoretical analysis of the fluorescence yield of rare earth ions in glasses containing small metallic particles,” Chem. Phys. Lett. 174(1), 13–18 (1990). [CrossRef]  

53. A. Archambault, F. Marquier, J.-J. Greffet, and C. Arnold, “Quantum theory of spontaneous and stimulated emission of surface plasmons,” Phys. Rev. B 82(3), 035411 (2010). [CrossRef]  

54. M. Finazzi and F. Ciccacci, “Plasmon-photon interaction in metal nanoparticles: Second-quantization perturbative approach,” Phys. Rev. B 86(3), 035428 (2012). [CrossRef]  

55. V. A. G. Rivera, F. A. Ferri, L. A. O. Nunes, A. R. Zanatta, and E. Marega Jr., “Focusing surface plasmons on Er3+ ions through gold planar plasmonic lenses,” Appl. Phys., A Mater. Sci. Process. 109(4), 1037–1041 (2012). [CrossRef]  

56. V. A. G. Rivera, Y. Ledemi, M. El-Amraoui, Y. Messaddeq, and E. Marega Jr., “Resonant near-infrared emission of Er3+ ions in plasmonic arrays of subwavelength square holes,” Proc. SPIE 8632, 863225 (2013). [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 Optical dark-field microscopic images of the NPs embedded in T5S12 (a) and T5G12 (b) samples.
Fig. 2
Fig. 2 Absorption spectra of (a) T5Sy, (b) T6Sy, (c) T5Gy, and (d) T6Gy glass samples. The inset shows a zoom of the absorption bands of transitions (a) 3H63F4, (b) 3H63H5, (c) 3H63H4 and (d) 3H63F2-3 in Tm3+ ions. All spectra show a high absorption of the Er3+ ions without NPs in comparison with the other samples, since the NPs increase the absorption these glasses.
Fig. 3
Fig. 3 Near infrared luminescence spectra under excitation at 976 nm of (a) T5Sy, (b) T6Sy, (c) T5Gy, and (d) T6Gy glass samples showing the 4I13/24I15/2 and 3F43H6 radiative transitions of the Er3+ and Tm3+ ions, respectively, and their splitting due to the Stark effect.
Fig. 4
Fig. 4 Schematic representation of the Er3+-Tm3+:NP interactions in two different scenarios: (a) the RE ions do not interact with each other, but with the NP and; (b) the interaction occurs between NP:RE and RE1:RE2. Equipotential surface (electric multipole coupling) with electric potentials −V/2 and + V/2 between NP:REs, d = 1,2 | r i R j | .
Fig. 5
Fig. 5 (a) Intensity of the emission peaks at 1535 ± 6 nm (4I13/24I15/2 transition of Er3+ ion) and 1795 ± 8 nm (3F43H6 transition of Tm3+ ion) for the TxMy samples studied. (b) Bandwidth of the emission band centered at 1535 ± 6 nm for the TxMy samples studied. (c) Emission spectra of the T5S08 and T6S10 samples which exhibit the most intense emission, and T5G08, T5G10, T6G02 and T6G12 samples which exhibit the broadest band.
Fig. 6
Fig. 6 Schematic energy level diagrams of (a) TxSy samples and (b) TxGy samples. Ground state absorption (GSA) of the Er3+ under 976 nm excitation resulting in a population inversion of the high (via ESA) and lower levels through nonradiative (NR) decays. Radiative ET among Er3+-Tm3+ ions (ETiE-iT), nonradiative energy transfer ETNR, and ET between the NP:RE (ETSNP or ETGNP). τNP is the finite plasmon lifetime (ns order) and the red curved lines represent the displacement of a small volume charge n j = q p,j ( r ) e i ω p t to the quantum system.
Fig. 7
Fig. 7 Lifetime of the Er3+:4I13/2 level for all samples studied. Dashed and dotted lines indicate, respectively, the lifetime of the sample without SNP and GNP for comparison. The T5G08, T5G10, T6G02 and T6G12 samples exhibit an expanded broadband (↔). The T5S08, T6S10 and T5G10 samples exhibit an improved luminescence (↑).

Tables (1)

Tables Icon

Table 1 Measured and calculated physical properties of TxMy glass samples: glass density, number (ionic or atomic) density N and average distance between rare-earth-ions and metallic-atoms (M0).

Equations (2)

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

r Er 3+ Ag 0 < r Er 3+ Au 0 for all the samples
r Tm 3+ Au 0 < r Tm 3+ Ag 0 for all the samples
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.