Maria S.
Batista
a,
Miguel P.
Dias
a,
Maria B.
Candeias
a,
Gabriel
Marques
a,
José D.
Gouveia
b,
Ana V.
Girão
c,
Florinda M.
Costa
a,
Joaquim P.
Leitão
a,
Luís
Rino
a,
Jonas
Deuermeier
d,
Elvira
Fortunato
d,
Rodrigo
Martins
d,
Ana
Pimentel
d,
Joana
Rodrigues
a,
Teresa
Monteiro
*a and
Sónia O.
Pereira
*a
ai3N, Department of Physics, University of Aveiro, 3810-193 Aveiro, Portugal. E-mail: sonia.pereira@ua.pt; tita@ua.pt
bCICECO, Aveiro Institute of Materials, Department of Physics, University of Aveiro, 3810-193 Aveiro, Portugal
cCICECO, Aveiro Institute of Materials, Department of Materials and Ceramic Engineering, University of Aveiro, 3810-193 Aveiro, Portugal
di3N/CENIMAT, Department of Materials Science, NOVA School of Science and Technology, Campus de Caparica, 2829-516 Caparica, Portugal
First published on 20th August 2025
Zn2GeO4 (ZGeO) phosphor exhibits a wide bandgap energy, making it highly suitable for luminescence-based applications spanning the entire electromagnetic spectrum, from ultraviolet (UV) to near-infrared (NIR) wavelengths. In this work, nominally undoped and Cr-doped ZGeO (ZGeO:Cr) were prepared by solid-state reaction. X-ray diffraction and Raman spectroscopy confirmed the monophasic willemite crystalline structure of the ZGeO and ZGeO:Cr samples, and X-ray photoelectron spectroscopy corroborated the identification of Zn, Ge, O, and Cr elements. X-ray photoemission indicates an insulator character for ZGeO and ZGeO:Cr. The oxide host revealed a direct bandgap energy of 4.77 eV, assessed by room temperature absorption, in line with the density functional theory (DFT) calculations that predicted 4.8 eV at the Γ point of the first Brillouin zone. Cr4+ was found to occupy distorted tetracoordinated Ge4+ sites with C1 symmetry, in agreement with the measured unfolded 3A2 → 3T1, 3T2 intraionic absorption. Er3+ and Mn2+ trace impurities occupy distorted Zn2+ sites, also with C1 symmetry. A Mn2+–O2− charge transfer state, placed 0.8 eV below the conduction band minimum, was identified by absorption measurements. In addition, as calculated by DFT, Cr3+ impurities exhibit lower energy formation when placed in distorted interstitial octahedral Zn–Ge and Zn–Zn rings with C3 and S6 symmetry, respectively. The identified site locations were found to be compatible with the measured unfolded 4A2 → 4T2, 4T1 intraionic absorption, with the ions subject to intermediate/low crystalline field strengths. The emission of the samples is dominated by structureless broad bands spanning from the visible to mid-infrared. A broad bluish-white emission results from an overlap of emitting centers related to both intrinsic defects (generating a blue emission) and trace impurities of Mn2+ (generating a green emission). Besides, Cr3+ and Cr4+ were found to coexist in the oxide host, and their multiple-site occupation is responsible for the observed broad emission bands in the NIR-I (700–950 nm) and NIR-II (1000–1700 nm) spectral regions, opening the way to the exploitation of the NIR luminescence for light-based devices.
With a room temperature (RT) bandgap energy reported between 4.4 eV and 4.9 eV,17,37–42 ZGeO exhibits a willemite-type crystalline structure with a rhombohedral unit cell (space group R, No. 148), with Ge and Zn ions occupying a Wyckoff position 18f,38,42–48 as is shown in Fig. 1. The crystalline cell is built by ZnO4 and GeO4 tetrahedra with cornered-shared oxygen, aligned parallel to the c-axis, and with lattice parameters a = b = 14.231 Å and c = 9.530 Å, α = β = 90° and γ = 120°.48–50 Zn2+ ions possess two inequivalent sites with reported average Zn–O bond lengths of 1.963 Å (Zn1) and 1.967 Å (Zn2). The remaining average bond length is 1.749 Å for Ge–O.20,47,50 Moreover, the lattice exhibits four-membered rings that encompass alternating corner-sharing ZnO4 and GeO4 tetrahedra and two types of six-membered rings: one consisting of corner-sharing Zn(1)O4 tetrahedra (radius of 196 pm) and the other one consisting of alternating corner-sharing Zn(2)O4 and GeO4 tetrahedra (radius of 277 pm).20,40,47–53
![]() | ||
Fig. 1 Structural model of zinc germanate obtained with VESTA using the CIF file: 1549041.54 |
Density functional theory (DFT) calculations predict that ZGeO has a direct bandgap, with the extremes of the conduction (CB) and valence (VB) bands occurring in the Γ point of the first Brillouin zone.38,53 The large energy separation between the CB and VB is prone to incorporating electronic energy levels inside the forbidden bandgap, generated either by native defects or by extrinsic dopants. J. Dolado et al.38 and S. K. Gupta et al.4 recently calculated the formation energies of intrinsic defects using DFT, showing that among the vacancy defects, the oxygen vacancy (VO) is the most favorable, while the germanium vacancy (VGe) is the least. In addition, interstitial zinc, Zni, in positions 1 (at the center of the Zn ring) and 2 (at the center of the Zn–Ge ring), exhibit the lowest formation energies, and antisite defects (ZnGe, GeZn) possess lower formation energy than the VO.4,38 Furthermore, the authors4,38 calculated the ZGeO electronic structure and intrinsic defect energy levels, establishing a correlation between the intrinsic defects and the luminescence observed in ZGeO. A schematic representation of the estimated energies for all the intrinsic defects obtained by the authors4,38 is shown in Fig. 2.
![]() | ||
Fig. 2 Schematic representation of the estimated energies for the ZGeO intrinsic defects based on the DFT calculations from ref. 4 and 38. |
Defects of intrinsic and extrinsic nature are typically responsible for the observed luminescence in undoped and intentionally doped ZGeO. Typically, using photons with energy equal to or higher than the oxide bandgap as excitation, the cryogenic and RT photoluminescence (PL) of the nominally undoped material is dominated by broad unstructured luminescence bands in the UV (UVL) and blue spectral regions.4,18,22,38,40,42,55–58 A UVL band peaked around 340–360 nm observed upon above bandgap excitation was recently reported on ZGeO microwires synthesized by a catalyst-free thermal evaporation method.38,58 The band was found to be an overlap of emitting centers, presenting either wide or narrow UVL bands, with the latter (peaked between 370–380 nm) revealing a two-component character and being also excited with below bandgap photon energies.38,58 Following the authors' DFT calculations,4,38 such high energy emission bands were attributed to radiative recombination involving electrons in either the VO or Zni donor levels with holes at the VB.38,58 In ZGeO, besides the UVL bands, a bluish-white band (BL) with a maximum located between 440 and 470 nm (depending on growth/synthesis parameters) was also ascribed to a recombination process involving intrinsic defects.4,18,22,40,55–57,59–61 The complex nature of this emission is still under debate in the literature. In particular, Z. Liu et al.40 proposed that the BL was related to a donor–acceptor pair (DAP) recombination mechanism also involving the VO and Zni as donors and ionized VGe and VZn as acceptors. After performing thermal annealing treatments in different atmospheres, Z. Gu et al.55 ascribed the recombination in this spectral region to a defect that involves the VO. J. Dolado et al.59 suggested that the VZn may be involved in the origin of the BL. In addition, the BL was considered as an overlap of several emitting centers, as proposed recently by M. Tinoco et al.57 and earlier by Li et al.60 These authors attributed the BL to germanium and oxygen-related defects. Moreover, the analysis of the recombination kinetics of the BL has been described as controversial in the literature. While Z. Liu et al.40 found a non-exponential decay behavior for the proposed DAP luminescence, M. Shang et al.18 reported a single exponential lifetime of 6.4 ns for the same emission, meaning that a complete model for the BL recombination still needs to be addressed.
The luminescence of ZGeO is highly sensitive to various factors, including synthesis/growth method, nature and purity of the precursors, as well as the temperature and atmospheric conditions during the production. In fact, in nominally undoped hosts or intentionally Mn-doped samples, the BL and an asymmetric broad green luminescence (GL) band with a maximum at ∼534 nm/2.32 eV have been observed4,20–22,38,41,46,48,51,55,59,61,62 when the material is excited with energies above the bandgap. The GL was also observed when ZGeO is excited either by high-energy electrons, resulting in green cathodoluminescence,37,63 or via carrier injection, enabling the observation of green electroluminescence.10,64 Despite the knowledge of the GL in ZGeO since at least the 1970s, its nature continues to be a matter of discussion.4,20,21,40,53
For instance, based on their DFT calculations, S. K. Gupta et al.4 recently reported that Zni in positions 1 and 2 are involved in the recombination processes of both the GL and BL emissions located at around 515 nm/2.41 eV. As mentioned above, besides the intrinsic defects, trace impurities (e.g., transition metal and/or lanthanide ions), even in nominally undoped samples, can originate optically active defects with electronic energy levels inside the material bandgap, resulting in luminescence bands over a wide range of the electromagnetic spectrum and with different spectral shapes. In the case of ZGeO, there is an enlarged consensus on the attribution of the ∼534 nm/2.32 eV GL band to the 4T1(4G) → 6A1(6S) transition of Mn2+ (3d5) impurities in substitutional Zn2+ sites.18,20–22,37,41,47,56,61,63,65
Electron paramagnetic resonance (EPR) measurements confirmed that Mn2+ occupies the two inequivalent Zn2+ sites (with one of the sites slightly preferred over the other one), as expected from the similarity of charge and ionic radius (0.60 Å and 0.66 Å for Zn2+ and Mn2+ in tetrahedral coordination, respectively).20,21,26,48 On the other hand, the low energy shift of the Mn2+ GL observed with increasing the Mn/Zn content, as reported by Feldman et al.,65 may be attributed to variations in crystal field strength arising from the distinct local environments of the two Zn sites.
Additionally, spin exchange interactions between Mn2+–Mn2+ pairs,22,66 as was previously observed in Mn-doped Zn2SiO4, which is isostructural with ZGeO43 may also contribute to this shift. Furthermore, although some authors37 have suggested that an emission band in the orange/red region (with maximum ca. 650 nm/1.91 eV) could be due to the presence of Mn4+ in an octahedral position in the ZGeO lattice, J. Q. Hu et al.50 recently reported that in ZGeO:Mn samples there is a reduction of Mn4+ to Mn2+ due to charge compensation effects.
As aforementioned, WBG oxides doped with chromium ions are of particular interest for the development of NIR pc-LEDs.26–36,67,68 For example, Cr4+ (3d2), which tends to occupy tetrahedral sites, is known to exhibit broad absorption bands in the spectral region where the main commercially available pump lasers operate, as is, for instance, the case of the chromium ion in Zn2SiO4, Mg2SiO4, and Ca2SiO4 crystals.26,27,69–77 In these oxides, when Cr4+ replaces Si4+ in distorted tetrahedral sites, strong intraionic polarized absorption bands take place between 500 nm and 800 nm and from 800 nm to 1200 nm due to the spin-allowed 3A2 → 3T1 (3F), 3T2 (3F) transitions,26,27,69–77 respectively. At RT, the broad emission of Cr4+ due to the 3T2 → 3A2 transition typically occurs in the NIR-II spectral region (1000–1700 nm), with a maximum ranging from 1000 nm to 1500 nm, depending on the crystal field strength.26,27,69–77 As mentioned, due to their broad pump absorption bands, Cr-ions have been extensively used as laser active media in several hosts for tunable solid-state lasers,75 and more recently, Cr-doped oxides are at the forefront of device development for applications in the NIR-I (700 to 950 nm) and NIR-II spectral regions.26,34,36,78,79
Considering the above, it is of great interest to explore the role of Cr ions in the ZGeO host, which, to our knowledge, has not yet been properly addressed. To the best of our knowledge, no red/NIR luminescence of Cr-activated ZGeO has been reported to date. As for the case of Mn2+ in Zn2+ sites, Cr4+ and Ge4+ exhibit the same valence and similar ionic radii in fourfold coordination (0.41 Å and 0.39 Å, respectively27) and, therefore, in Cr-doped ZGeO, it is expected that Cr4+ in Ge4+ sites could be optically activated, as will be further discussed in this work. On the other hand, Cr3+ (3d3) has an ionic radius of 0.46 Å in four-fold coordination,68 and 0.61 Å in a six-fold coordination (close to the Zn2+ in four-fold coordination), which, according to Y. Cong et al.,56 is expected to substitute Ge4+ in ZGeO. According to the Tanabe and Sugano diagrams80 for this electronic configuration, a 4A2(4F) ground state is expected for the ion in octahedral coordination and, depending on the crystal field strength, the first excited state is either the 2E(2G) (strong fields) or 4T2(4F) (weaker fields). Therefore, the luminescence from the first excited level to the ground state is mainly characterized by narrow emission lines, the well-established R-lines, due to the 2E → 4A2 transition, or by a broad emission band due to the 4T2 → 4A2 transition.30,31,74,76,77,81,82 In addition, 4T2 → 4T1 NIR-I luminescence due to Cr3+ in a tetrahedral coordination environment has also been reported in several oxide hosts, in which the ion experiences weaker crystal field strengths than those in octahedral sites.68,82–84 However, as mentioned by Shang et al.,68 Jahn–Teller effect on the ground state and/or inherent lattice distortion enhances the nonradiative transition probability between the aforementioned electronic states, which may be the origin of the literature ambiguities concerning Cr3+ luminescence on Td site symmetry, with widely accepted Cr3+ activators tending to occupy Oh (or near) lattice sites.68,85
In this work, nominally undoped (with Mn and Cr as trace impurities) and intentionally Cr-doped ZGeO were synthesized by conventional high-temperature solid-state reaction, and an in-depth optical characterization of these oxide materials was performed. Apart from the BL and GL bands, unreported emission bands in the NIR-I and NIR-II spectral regions were observed. Based on the absorption, temperature, and excitation energy dependence on luminescence, the emission bands were assigned to Cr3+ in distinct distorted octahedral sites and to Cr4+ in tetrahedral Ge4+ sites, respectively, which agrees well with the performed structural and elemental analysis. The experimental evidence is supported by first-principles calculations on the formation energies of Cr in ZGeO, their site occupation, and valence states. Moreover, from photoluminescence excitation (PLE) measurements, the preferential population mechanisms of the emitting centers were established, highlighting the role of intrinsic defects, Mn2+–O2− charge transfer state (CTS), and Cr3+/Cr4+ excitation bands. Therefore, the present theoretical and experimental investigation elucidates the mechanisms of the broad emission bands covering the visible, NIR-I, and NIR-II spectral regions in ZGeO:Mn,Cr materials, opening the way to future developments of this emerging oxide host for luminescence-based applications.
For the morphological and elemental composition characterization, the scanning electron microscopy (SEM) images were obtained using a field-emission Schottky-gun (FEG-SEM) SEM Hitachi SU70, operated at 15 kV, and equipped with an AXS Bruker energy dispersive X-ray fluorescence (EDX) detector. Additionally, for the identification of trace impurities, ICP-OES analysis was performed using a Horiba JobinYvon Activa M instrument.
X-ray photoelectron spectroscopy (XPS) measurements were carried out using a Kratos Axis Supra instrument with monochromated Al Kα radiation (1486.6 eV). Detailed scans were obtained at an X-ray power of 150 W and a pass energy of 10 eV. Electron flood gun charge neutralization was used, and all peaks were referenced to C 1s at 284.8 eV. Data analysis was conducted with CasaXPS software.
Diffuse reflectance (DR) of the undoped and intentionally doped pellets was measured at RT in the wavelength range 190–1600 nm using a UV/Visible GBC Cintra-303 equipment in the absorbance mode, with a bandwidth of 2 nm, a speed velocity of 100 nm min−1, and a data pitch of approximately 0.520.
Fourier-transform infrared (FTIR) absorption measurements were conducted in an FTIR Bruker Tensor 27 using a diamond ATR Golden Gate, in absorbance mode, with a resolution of 2 cm−1 and 256 scans in the range 4000–400 cm−1.
The RT steady-state PL and PLE data were obtained in a Fluorolog®-3 from Horiba Scientific. A continuous Xe lamp of 450 W was used as the excitation source, coupled with an excitation monochromator, Gemini 180, with a diffraction grating of 1200 grooves mm−1 and a focal distance of 180 mm. The scanning emission monochromator is an iHR550 with a diffraction grating of 1200 grooves mm−1 and a focal distance of 550 mm. The detector used is an R928P PMT, which is sensitive to wavelengths in a range of 200 to 850 nm. For the RT time-resolved PL (TRPL) measurements, the same Fluorolog®-3 system was used, in this case with the aid of a pulsed Xe lamp (operating at up to 25 Hz, pulse lamp width of 3 μs) coupled to the same monochromator. The measurement conditions were set to variable windows (duration of signal acquisition) of 1, 2, and 5 ms, and several time delays after flash were employed (from 0.08 to 30 ms). For the lifetime measurements, the samples were excited with appropriate excitation wavelengths to maximize the PL signal, and the time evolution of the PL intensity was assessed.
For the RT and low temperature PL data acquisition in the UV-visible range, a dispersive system SPEX 1074 Czerny–Turner monochromator with a diffraction grating of 1200 grooves mm−1 and a focal distance of 1 m was used. This equipment was coupled to a RCA C31034 photomultiplier detector. The excitation source used was the 325 nm line of a cw He–Cd laser. The samples were placed in a cold finger of a closed-cycle He cryostat under vacuum conditions, which allowed the temperature to reach 15 K. Temperature-dependent PL spectra were performed between 15 K and the RT. The spectra acquired in both systems were corrected for the optics of the system and the detector response. For quantitative analysis, the recorded data were corrected for the Jacobian transformation.
Vis-NIR/MIR PL measurements were also carried out in the range of 22000–5000 cm−1 (455–2000 nm) using a Bruker Vertex 80v Fourier Transform Infrared (FTIR) spectrometer equipped with InGaAs and Si detectors. To remove the alignment laser line at 632.8 nm, a long-pass optical filter with a cut-off wavelength of 700 nm was used. The samples were inserted and mounted on an OptistatCF (Oxford Instruments) cryostat, and the spectra were measured from 70 K to RT. As excitation sources, the 457.9 nm (DPSS (CVI Melles Griot)), 532 nm (MGL-F-532 (CNI)) and 808 nm (Roithener LaserTechnik, RLDB808-500-5) laser lines were used.
The simulations were run on a Zn2GeO4 unit cell, with periodic boundary conditions in all cartesian directions. The explicitly treated electrons were 3d10 4s2 for Zn, 3d10 4s2 4p2 for Ge, 2s2 2p4 for O, 3p6 3d5 4s1 for Cr, 5p6 5d1 6s2 for Er, and 3p6 3d6 4s1 for Mn, with the remaining electrons included in the frozen core and tackled using the projector augmented-wave method.88 The Brillouin zone was sampled using only the Gamma k-point, due to the extreme computational cost of hybrid DFT calculations when using the plane-wave formalism.89 The plane-wave basis had a cutoff value of 415 eV, and the force (energy) convergence criterion for structural (electronic) relaxation was 0.01 eV Å−1 (10−5 eV).
Zn2GeO4 has S6 symmetry and the following lattice sites: all Ge4+ sites are equivalent and tetrahedral (GeO4); there are two inequivalent Zn2+ sites, both tetrahedral (ZnO4); and there are four inequivalent O2− sites, all trigonal planar, with O bonded to two Zn2+ and one Ge4+.
The crystallinity of the samples was also assessed by Raman spectroscopy, as shown in Fig. 3b. The spectra of ZGeO and ZGeO:Cr are similar, with the vibrational modes located at the same frequency with almost the same relative intensity and full-width half maximum (FWHM). The strongest resonance occurs at 802 cm−1 (A(2)g) and has been assigned to the stretching vibration O-Ge–O bonds in GeO4 tetrahedra.42,58,62,90–93 Besides the observation of a vibrational mode at 861 cm−1, two other maxima were measured at 746 cm−1 (A(1)g) and 779° cm−1 (E(4)g), which are associated with Ge–O–Zn symmetric and asymmetric vibrations, respectively.42,58,62,90–93 The spectral shape of this peak is compatible with the presence of another peak at 751 cm−1 (E(3)g) which, although unresolved, cannot be definitively ruled out. The low-frequency modes (<600 cm−1) corresponding to O–Zn–O and Zn–O–Ge bending and stretching vibrations91,93 were also identified in the studied samples.
The vibrational host absorption is also shown in Fig. 3c, which presents the FTIR absorption spectra for both undoped and doped samples, in which eight modes were identified. In line with what has been reported in the literature,22,93,94 the absorption bands at 450–650 cm−1 and 700–900 cm−1 were assigned to the vibrational modes of ZnO4 and GeO4 tetrahedra, respectively.
Fig. 3d and e show SEM images of the ZGeO and ZGeO:Cr samples, respectively, with the same magnification. The microscopy images display a similar cohesive and dense morphology for both samples. Additionally, Fig. 3f shows the EDX spectra with the corresponding indexation of elements. EDX analysis indicates the presence and homogeneous distribution of the constituent chemical elements, Zn, Ge, and O. The Cr element was also identified in the doped sample, as were impurities of other elements such as Si, Al, and Ca, which arise from the accessories used for SEM-EDX observation.
A more detailed chemical analysis of the phosphors was performed by XPS. The survey XPS spectra (Fig. 4a) exhibit the presence of Ge, Zn, O, and C in both samples; additionally, in the doped sample, Cr was also identified. The spectra were corrected to adventitious carbon (C 1s) at 284.8 eV. For the assessment of the XPS detailed spectra of the elements, Fig. 4b–d, a Shirley background was applied for fitting purposes.
Fig. 4b shows the O 1s spectrum. The emission was deconvoluted into two peaks at 530.6 and 531.4 eV for ZGeO, as well as 530.9 and 531.9 eV for ZGeO:Cr, which were assigned to lattice oxygen and surface-adsorbed oxygen.90,95,96 The observed increase in the binding energies from the non-doped sample to the doped one can be related to the electron charge compensation97 due to the introduction of chromium in the matrix, for instance, when Cr4+ replaces a Ge4+. The fine XPS spectra (Fig. 4c) show one of the two Zn 2p peaks characteristic of ZGeO and ZGeO:Cr at 1021.1 and 1021.4 eV, which are in agreement with the ones reported previously.90,95 Moreover, the second Zn 2p peak was identified in both samples (data not shown), with an energy separation of approximately 23.1 eV between the two Zn 2p peaks, consistent with previously reported values for this host material.90,95 The fine XPS spectra displayed in Fig. 4d reveal the presence of the Ge 3d peak located at 31.6 eV and 32.0 eV for ZGeO and ZGeO:Cr, respectively, which are close to the values reported elsewhere for this host material.90,95,98
In the undoped sample, no evidence of Cr was found. However, as expected, in the ZGeO:Cr sample, the Cr 3p was identified at circa 43.5 eV, confirming the presence of this dopant (Fig. 4e).99 The region around 590 to 575 eV, corresponding to Cr 2p (2p1/2 and 2p3/2 orbitals of chromium), has been used to identify the presence of trivalent and tetravalent states of chromium in similar phosphors (without Zn), such as Mg14Ge5O24:Cr3+, Cr4+ (ref. 100) and BaTiO3:Cr3+, Cr4+.96 However, Zn LMM Auger also occurs at this binding energy,101 so an accurate confirmation of the valence state of chromium is hindered for ZGeO and ZGeO:Cr. Nevertheless, as will be discussed later, Cr3+ and Cr4+, as well as Mn2+ (not detected by XPS), were identified by optical measurements in undoped and doped samples.
To quantify the elements present in each sample, the Cr 3p emission was used, but as it is shallow, a custom polynomial background correction was applied in this case, intended to mimic the background of the undoped sample. Additionally, the Zn 3p emission was used due to its kinetic energy being closer to that of the other cations than in the case of Zn 2p. The results can be found in Table 1, where a ratio of Zn:
Ge of 2
:
1 for the ZGeO sample was determined, whereas the doped sample showed a zinc deficiency (Zn
:
Ge = 1.82
:
1) and a chromium fraction of 0.01 related to Ge, which is in agreement with the nominal percentage of Cr used for sample preparation.
Name | Position (eV) | ZGeO | ZGeO![]() ![]() |
x/Ge | Area | At% | x/Ge |
---|---|---|---|---|---|---|---|
Area | At% | ||||||
a The C percentage was removed to determine the ratio between an element (x) and Ge. | |||||||
O 1s | 531.0 | 16![]() |
40.83 | 4.81 | 16![]() |
47.42 | 4.35 |
C 1s | 285.0 | 4767.34 | 33.47 | 2781.51 | 21.78 | ||
Ge 3p | 124.0 | 4854.36 | 8.48 | 1 | 5589.31 | 10.89 | 1 |
Zn 3p | 89.0 | 8282.86 | 17.22 | 2.03 | 8525.18 | 19.77 | 1.82 |
Cr 3p | 43.5 | — | — | — | 17.61 | 0.14 | 0.01 |
![]() | ||
Fig. 5 (a) RT normalized absorption spectra of ZGeO and ZGeO:Cr samples. (b)–(d) adapted Tanabe and Sugano diagrams for the Cr4+ 3d2 electronic configuration in Td site symmetry, Cr3+ 3d3 electronic configuration in Oh site symmetry (for simplicity, the gerade (g) sub-indices were omitted), and Mn2+ 3d5 electronic configuration in Oh site symmetry, respectively. (b) to (d) Adapted from ref. 104 and 105. |
For the ZGeO lattice, and due to the charge and size of the Cr4+ that closely match that of Ge4+, it is expected that Cr4+ will enter the lattice by substituting the fourfold coordinated germanium cations. In this valence state, and by the Tanabe and Sugano diagram for the 3d2 electron configuration in tetrahedral symmetry (Fig. 5b), the ground 3F free ion term splits, in increasing energies, into 3A2, 3T2, and 3T1 electronic states, while the 1D term is split into the 1E, 1T2 levels, and the 3P excited state into the 3T1 level.104 Accordingly, spin and electric dipole-allowed transitions are expected to be observed from the 3A2 electronic ground state to the 3T1 (3F) and 3T1 (3P) excited levels, and if the relaxation of the selection rules occurs due to symmetry lowering, the 3A2 →3T2 (3F) absorption can also be observed to some extent at lower energies. This is, for instance, the case corresponding to the ion in a distorted tetrahedral environment (e.g., a trigonal distortion (C3v) followed by a Cs component), which is known to split the triplet states into three orbital components. Such symmetry lowering results in pronounced polarized absorption transitions as were already reported for the case of the Mg2SiO4 and Ca2GeO4 forsterite and cunyite olivine crystals, in which the 3A2 →3T2 (3F) absorption transitions are seen to occur between 800 nm/1.55 eV and 1100 nm/1.13 eV and the 3A2 →3T1 (3F) between 500 nm/2.48 eV and 800 nm/1.55 eV.71–73,106,107
Theoretical calculations performed on the Cr4+ site location on ZGeO support a similar interpretation. Indeed, although we mentioned that all Ge sites are equivalent and tetrahedral (GeO4), these are not regular tetrahedra (with Td symmetry). The Ge–O bond lengths and angles are not all equal, reducing the symmetry of Ge sites to the lowest one, C1. Therefore, any substitutional atom at a Ge site does not see a local Td symmetry, but C1 instead, even before any lattice distortions occur because of the substitution by a distinct element. According to our DFT calculations, replacing a Ge4+ with a Cr4+ causes nearly no lattice strain, indicating very high structural stability of this defect, leading to a fundamental electronic state placed at 1.25 eV above the VB maximum (Fig. S1 of SI). Two of the bond lengths with the neighboring O atoms do not change (up to 0.01 Å), the third one increases by 0.01 Å, and the fourth decreases by 0.01 Å (Table 2), so the average Cr–O distance is the same as the average Ge–O distance.
Ge4+–O (Å) | Cr4+–O (Å) |
---|---|
1.76 | 1.75 |
1.76 | 1.76 |
1.76 | 1.76 |
1.76 | 1.77 |
Therefore, from the spectral shape of the unpolarized RT absorption spectra shown in Fig. 5a, and by analogy with the above-mentioned behavior of the chromium ions in the tetravalent charge state in other hosts, we attribute the longer wavelength absorption bands to Cr4+ in distorted tetrahedral sites on the ZGeO lattice. The bands at 636 nm/1.95 eV, 694 nm/1.79 eV, and 767 nm/1.62 eV correspond to the split 3A2 → 3T1 (3F) transition and the weaker bands at 898 nm/1.38 eV and 1129 nm/1.1 eV are due to the 3A2 → 3T2 (3F) transition of Cr4+ in C1 lower site symmetry.
For the case of the intentionally Cr-doped sample (ZGeO:Cr), besides the absorption of the Cr4+, an increase in the relative intensity of the absorption bands located at shorter wavelengths (between 360 nm and 640 nm) is noticed, which are likely to be related to the introduced dopant. A careful inspection allows the identification of the absorption bands already present in the ZGeO sample, in which chromium was found as a trace impurity. Considering this fact, the absorption bands in this spectral region were attributed to the trivalent chromium in the ZGeO host. In this charge state, Cr3+ is usually found in octahedral coordination in several hosts, which is energetically favorable.74,100,108,109 To evaluate the preferred ion site location in the ZGeO host, we performed theoretical calculations, also considering the possibility of the ion in four-fold coordination. As in the case of Ge, the ZnO4 tetrahedra are not regular. There are two inequivalent Zn sites. In one of them, the calculated Zn–O bond lengths range from 1.95 to 1.97 Å, whereas in the second one, these lengths are between 1.95 and 1.99 Å. As before, a substitutional Cr3+ at a Ge or Zn site does not see a local Td symmetry, but C1 instead. Moreover, none of the substitutional sites of the Zn2GeO4 lattice is octahedral. However, it is possible for Cr3+ to bind to six O atoms in a distorted octahedron configuration: in one of the two interstitial positions, at the center of a Zn–Ge ring, or of a Zn ring. The tetrahedral sites are almost regularly tetrahedral. As mentioned, the bond lengths are the same up to 0.01 Å. The perfect tetrahedral angle is 109.5°, and the O–Ge–O angles in Zn2GeO4 have values very close to this (Fig. 6a). For the octahedral sites, two possibilities must be considered: (i) the center of a Zn–Ge ring and (ii) the center of a Zn ring. Fig. 6b shows a Cr3+ at the interstitial site at the center of a Zn–Ge ring (C3 symmetry), viewed along the c-direction. All bond lengths are equal (2.16 Å) up to differences of 0.01 Å, but the O–Cr–O bond angles are not 90° and 180°, as in a regular octahedron. In a regular octahedron, all the angles would be 90°, as shown in Fig. 6b. Instead, they are all lower than that, so this corresponds to a squeezed octahedron, with a side view represented in Fig. 6c. A similar conclusion can be drawn for the case of an interstitial Cr3+ at the center of the Zn ring (S6 symmetry), as shown in Fig. 6d, where all bond lengths are similar (2.14 Å), but the bond angles are 75.9°.
![]() | ||
Fig. 6 (a) GeO4 tetrahedral site in ZGeO with bond lengths mentioned in Table 2 and O–Ge–O close to the 109.5° expected for a perfect tetrahedron. (b) Interstitial Cr3+ site at the center of a Zn–Ge ring (C3 symmetry), viewed along the c-direction. (c) Side view of the octahedron shown in (b). (d) Interstitial Cr3+ site at the center of a Zn ring (S6 symmetry), viewed along the c-direction. |
To find the thermodynamically preferred site of Cr3+, we calculated its formation energy on each site, Eform, using the following expression:89,110
Eform = Edefect − Epristine + Eremoved − Eadded + q(EVBM + EFermi) | (1) |
Hence, as the investigated samples evidenced a substantial isolating behavior as assessed by XPS measurements (Fig. S2), the Fermi level energy should be placed near the middle of the bandgap energy, meaning that Cr3+ ions are likely to preferentially occupy the two distorted octahedral interstitial sites at Zn–Ge and Zn rings, adding occupied states near 0.6 eV above the VB maximum and unoccupied states near the CB minimum (Fig. S1). A similar trend for interstitial Cr3+ was recently reported for the case of isostructural Zn2SiO4.108 The free ion terms of the Cr3+ (3d3) split in the cubic field are illustrated by the Tanabe and Sugano diagram of Fig. 5c. The 4F ground state unfolds into the 4A2, 4T2, and 4T1 electronic states, the excited 4P term to the 4T1, and the 2G state to the 2E, 2T1, 2T2 and 2A1. Under this site symmetry, Cr3+ is known to give rise to strong absorption bands in the visible spectral range in several hosts,74,100,108,109 mainly due to the spin-allowed and parity-forbidden 4A2 → 4T2, 4T1 (4F) transitions. The spin and parity forbidden 4A2 → 2E, 2T1 (2G) may appear as very narrow absorption lines overlapped with the 4A2 → 4T2 (4F) absorption band.105 From Fig. 5a, an increase in the relative intensity of the visible and UV absorption peaked at 584 nm/2.12 eV, 455 nm/2.73 eV, and 415 nm/3.0 eV is noticeable, which corresponds to the typical spectral region of the 4A2 → 4T2, 4T1 (4F) transitions of Cr3+ ions in intermediate crystal field values. Therefore, and by comparison with the theoretical calculations, we assign the high energy absorption bands to the trivalent chromium in octahedral distorted sites with C3 and S6 symmetry in the ZGeO lattice. It should be emphasized that the low energy Cr3+ absorption transition partially overlaps with the absorption of Cr4+ on distorted tetrahedral sites (C1). As a first conclusion, we can confirm that the measured RT absorption is consistent with the simultaneous presence of the trivalent and tetravalent chromium charge states in the ZGeO lattice.
From the precedent analysis, the average crystal field strength (Dq) felt by chromium ions in the tetra and trivalent charge states, when located in a distorted tetrahedral or octahedral environment, can be estimated.104,111 Considering the mean energy value of the 3A2 → 3T2, 3T1 (3F) transitions of the Cr4+ and 4A2 → 4T2, 4T1 (4F) transitions of Cr3+, and the corresponding energy differences,104,111Dq/B values of ∼2.34 and ∼2.44 were estimated for Cr4+ and Cr3+, with B Racah parameters of ∼475 cm−1 and ∼700 cm−1, respectively. These average values suggest that the intermediate crystal field strength of Cr3+ ions is close to the crossing point between the 4T2 (4F) and 2E (2G) levels, impacting the expected observed luminescence, as will be further discussed.
Regarding the absorption band at 313 nm/3.96 eV, its corresponding signature is less evident than the previous ones, and several hypotheses should be considered. From the estimated Dq/B values for Cr4+, and considering the Tanabe and Sugano diagram (Fig. 5b), a transition from the ground 3A2 (3F) to the excited 3T1 (3P) state can be excluded, as a high Dq/B value (∼3) would be required. In addition, and using a similar argument, a lower Dq/B value (∼2) would be necessary to observe the 4A2 (4F) → 4T1 (4P) transition of Cr3+. Therefore, it is unlikely that the absorption band at 313 nm/3.96 eV could be assigned to internal transitions of the Cr4+ or Cr3+. As stated in the introductory section, a wide variety of defects are expected to be present in both samples, as is the case of intrinsic defects and other trace impurities. As mentioned before, DFT calculations4,38 place one of the VO energy levels close to the energies of the observed absorption band. Therefore, we cannot exclude that the 313 nm/3.96 eV absorption band could be due to such a defect. On the other hand, the presence of Mn2+ (3d5) as a common trace impurity in ZGeO could also lead to absorption bands at the mentioned higher energies, as discussed in the next section.
![]() | ||
Fig. 8 RT absorption and PLE spectra monitored at the BL, GL, and NIR-I band maxima for (a) ZGeO and (b) ZGeO:Cr samples. (c) RT and 70 K PL spectra obtained with above and below bandgap excitation for the ZGeO and (d) ZGeO:Cr samples. (e) Comparison between the steady-state and time-resolved PL for the ZGeO sample, as described in the text. (f) Rise and decay times for the BL and GL observed in the undoped and Cr-doped samples acquired with different photon excitation (symbols: experimental data; solid lines: fits to eqn (2)). All spectra are normalized to the intensity maximum. |
In addition, the Mn2+ GL is also populated by lower energy photons, namely via the excitation band centered at ∼313 nm/3.96 eV, which coincides with the unassigned defect absorption band mentioned above when discussing Fig. 5a. Although this absorption band is close in energy to one of the theoretically DFT-predicted oxygen vacancy donor energy levels,38 a more likely explanation is that the absorption band arises from the Mn2+–O2− CTS, as reported by others,22,41,46,50 with the CTS energy level placed ca. 0.8 eV below the conduction band. Therefore, and as mentioned, for observing the Mn2+ GL, two main population paths should be considered: (i) pumping the samples with above bandgap energy photons, and (ii) exciting the samples with subgap energy photons coincident with the CTS absorption.
The interplay between the two overlapped emitting centers (BL and Mn2+ GL) can be further visualized in Fig. 9 and in the multimedia file in the SI. Here, upon continuous band-to-band excitation, the RT Mn2+ GL of ZGeO can be seen with the naked eye for a few seconds, after which the BL recombination prevails. A similar observation was also recently reported by Yue et al., who assumed both the BL and GL as self-activated DAP recombination involving intrinsic defects.61 The identified behavior indicates that carrier trapping/de-trapping mechanisms and distinct feeding times of the defect emitting levels from which the GL and BL originate should be present in the studied samples. Such findings are corroborated by the measured transient behavior of the luminescence at RT (Fig. 8f) upon band-to-band (260 nm) or sub-bandgap (310 nm) photon excitation. As seen, before the decay, a rise time of the BL and GL intensities occurs, meaning that distinct population pathways are responsible for feeding the BL and Mn2+ GL 4T1 emitting levels, respectively. The temporal dependence of the emission intensity I(t) shown in Fig. 8f was fitted considering that I(t) is proportional to the number of electrons in the emitting level,113 which, in our case, follows the approach described by V. Lojpur et al.,114
![]() | (2) |
![]() | ||
Fig. 9 Frames of the multimedia file showing the color shifting emission upon continuous band-to-band photon excitation (254 nm) for the undoped (left) and doped (right) samples at different times. |
Emission band | BL | GL | RL | |
---|---|---|---|---|
Sample | τ exc = 260 nm | τ exc = 310 nm | τ exc = 310 nm | |
ZGeO | τ r (ms) | 0.044 | 0.237 | — |
ZGeO:Cr | 0.014 | 0.335 | — | |
ZGeO | τ d1 (ms) | 14.6 | — | — |
ZGeO:Cr | 13.8 | — | — | |
ZGeO | τ d2 (ms) | 2.72 | 3.61 | — |
ZGeO:Cr | 1.67 | 3.36 | — | |
ZGeO | τ d3 (ms) | — | — | 1.25 |
ZGeO:Cr | — | — | 1.73 | |
ZGeO | τ d4 (ms) | — | — | 3.98 |
ZGeO:Cr | — | — | 4.45 |
When exciting the samples at the Mn2+–O2− CTS absorption band (310 nm/4.00 eV), a single exponential decay of ca. 3.6 ms is measured for the spin and parity forbidden 4T1 → 6A1 transition of Mn2+ GL for the ZGeO sample. This intrinsic radiative decay is slightly shortened when the band-to-band excitation is performed and monitored at 480 nm, suggesting that additional energy transfer processes may occur, as expected from the identification of spectral overlap between the GL emission and Cr ions absorption (Fig. 5a). In addition, and as expected, the BL presents a slower radiative decay of 14.6 ms, in line with the observed transient PL (Fig. 8e). Moreover, it is important to mention that the rise times measured for the BL and GL under the different excitation conditions differ by almost one order of magnitude. These findings allow us to hypothesize that, upon band-to-band excitation, nonradiative carrier relaxation takes place from the host to higher excited defect levels (e.g., Mn2+–O2− CTS and/or other defects), slowly populating the Mn2+ 4T1 emitting level, leading to the observed GL. In addition, competitive nonradiative processes through the crossover of the defect levels and/or energy transfer between other intrinsic/extrinsic defects should arise, explaining the de-trapping process that leads to the change of GL to BL under continuous band-to-band excitation, as shown in Fig. 9 and in the multimedia file in SI. From our data and following the DFT calculations performed by Gupta et al.4 and Dolado et al.,38 it is hypothesized that the mentioned electronic states crossover may involve other intrinsic defects, e.g. Zni, which are the native defects with lower formation energy, and the estimated location of their energy levels might justify the observation of the BL. Therefore, our data are consistent with a description of the broad bluish-white emission band as an overlap of the intraionic GL luminescence of Mn2+ in Zn2+ sites with the BL likely related to Zni intrinsic defects.
For the case of the doped ZGeO:Cr sample, as shown in Fig. 8b and d, both BL and Mn2+ GL are observed, although with distinct relative intensities compared with the nominally undoped ZGeO. In addition, and despite the overlapped emissions, for this sample, only the BL is seen with the naked eye upon continuous band-to-band excitation, as visualized in Fig. 9 (multimedia file in SI). Also, as shown in Fig. 8f, the luminescence intensity decay of the Mn2+ GL for the ZGeO:Cr sample is shortened when a comparison is made with the non-intentionally doped sample. As previously stated, energy transfer processes are expected to occur in line with the increase in the relative intensity of the Cr-related absorption, likely explaining the decrease in the measured lifetime. The same tendency is observed for the BL, showing that the populations of the GL and BL are strongly conditioned by trapping/de-trapping and energy transfer phenomena arising from the presence of the intrinsic and extrinsic defects.
Temperature-dependent PL measurements between 15 K and RT were also performed on ZGeO and ZGeO:Cr samples. With 325 nm photon excitation, which corresponds to an excitation on the Mn2+–O2− CTS defect absorption (Fig. 8a and b), the PL intensity of the intraionic 4T1 → 6A1 transition of Mn2+ trace impurities decreases with increasing temperature for both samples, with different thermal quenching factors, as shown in Fig. S4. In particular, 30% and 19% of its original intensity at 15 K were found at RT, for ZGeO and ZGeO:Cr samples, respectively. Under steady-state conditions, the quenching of the PL intensity due to any thermally activated process depends on the rate equations that describe the population and depopulation of the emitting excited state. Considering negligible nonradiative processes at low temperatures, the radiative quantum efficiency can be defined as η = krad/(krad − knrad) where krad,nrad are the radiative (krad = 1/τrad) and nonradiative (knrad = 1/τnrad) transition probabilities. Assuming a classical Mott law, the nonradiative transition probability can be written as knrad = Aexp(−Ea/kBT) and, therefore, the temperature dependence of the Mn2+ spectrally integrated GL intensity can be well accounted for,
I(T) = I0[1 + C![]() | (3) |
The peak position of the high energy emission band strongly depends on temperature, excitation energy, and Cr3+ amount (Fig. 10c). As The RL/NIR-I is preferentially excited by photons with an energy that matches the defect absorption band, i.e., the Mn2+–O2− CTS, as well as by intraionic Cr3+ absorption (Fig. 8a and b), the luminescence could likely be assigned to the Cr3+ ions in line with what is observed in other hosts, even knowing that the emission is strongly dependent on the crystal field strength.30,31,74,76,77,81,105,111,116–119 Typically, in octahedral coordination, the Tanabe and Sugano energy levels diagram for the 3d3 electron configuration (Fig. 5c) allows the identification of the first excited electronic state of Cr3+ as 2E or 4T2, for high or low crystal field strengths, respectively. In addition, Fig. 5c shows that the energy location of the 2E state is almost insensitive to an increase in the strength of the crystal field, while the energy of the 4T2 and 4T1 (4F) excited states increases with increasing Dq/B. As stated in the Introduction, in several hosts and when in octahedral sites, the intraionic Cr3+ luminescence to the 4A2 ground state is characterized by: (i) well-defined zero phonon lines (ZPL), the so-called R-lines, and by rich vibrational sidebands for the case of the 2E→4A2 transition, which preferentially occurs when the 2E state is placed at lower energies than the 4T2, i.e., when the ion is placed in sites with high crystalline field strengths;30,74,81,105,118,119 or (ii) by broadband recombination corresponding to the 4T2 →4A2 transition, usually observed when the 4T2 is at lower energy than the 2E, i.e., when the ion is under a low crystalline field strength environment.31,74,76,77,105,111,116,117 Furthermore, we note that from our DFT calculations, for an insulating ZGeO, Cr3+ is expected to occupy two interstitial sites. This suggests that at least two Cr3+ centers, located in distorted octahedral sites (C3 and S6 symmetries), may be optically activated. From the absorption measurements, we have identified that the ions are, on average, under an intermediate crystal field, meaning that the luminescence could arise from the components of the 2E and 4T2 excited states, upon symmetry lowering (a doublet and a singlet for the case of the triplet state in the case of a trigonal/rhombohedral distortion; or a fully degeneracy splitting in singlet states if a further orthorhombic distortion is considered). These assumptions allow us to explain the observed RL/NIR-I luminescence measured with two distinct spectroscopic systems (see experimental details) and with different excitation wavelengths, as shown in Fig. 8c, d and 10c (and Fig. S6). For the case of the ZGeO sample, and as displayed in Fig. 10c, the 70 K PL spectra evidence structureless broad emission bands with maxima at 812 nm/1.53 eV, 800 nm/1.55 eV, and a short wavelength shoulder at ca. 740 nm/1.67 eV when the excitation is performed with 325 nm, 457.9 nm, and 532 nm photons. The absence of lines/structure at this temperature may indicate that: (i) if the transitions are due to the 2E→4A2, the ZPLs are already quenched at 70 K; (ii) the emission could preferentially arise from the split 4T2→4A2, which usually gives rise to a broad unstructured band.
As the bands’ peak position depends on the excitation wavelength, one cannot discard the hypothesis that overlapped distinct Cr3+ emitting centers could occur. Therefore, for the ZGeO sample, in the 70 K spectra, we designated by Cr3+(1) the center with maximum at 812 nm/1.53 eV preferentially excited with 457.9 nm, and by Cr3+(2) the one with maximum at 800 nm/1.55 eV, that is 23 meV apart from the former and is preferentially observed with 532 nm photons. In addition, a third center can be distinguished for the case of the ZGeO:Cr sample (Cr3+(3)). The latter exhibits a similar peak position to Cr3+(1) and is detected with photon excitation at 325 nm, 457.9 nm, and 532 nm.
This site-selective wavelength excitation 70 K PL allows us to conclude that at least two optically active Cr3+ centers coexist in the ZGeO sample, in line with the theoretical DFT predictions. Moreover, the RL/NIR-I emission bands from the Cr3+ (1 to 3) centers are sensitive to the temperature rising from 70 K to RT. As seen in Fig. 10c (see also Fig. S6), a high-energy shift of the emission maxima with increasing temperature was identified, indicating that a thermal population of higher excited states occurs. Particularly, an energy separation of ca. 73 meV and 50 meV was estimated from the band maxima for the case of Cr3+(1) and Cr3+(2) in ZGeO sample, respectively, whereas a value of ca. 23 meV was found for the Cr3+(3) center in the ZGeO:Cr sample. These results also evidence that Cr3+(2) and Cr3+(3) are not the same center, despite their similar spectral shape at 70 K.
At longer wavelengths (Fig. 10d), and still in the NIR-I region, two main bands are observed centered at 880 nm/1.41 eV and 940 nm/1.32 eV. These emission bands are detected upon excitation on the Cr3+ absorption (e.g., 458 nm and 532 nm) and are absent for longer wavelength photon excitation (808 nm), in which the Cr4+ absorption takes place (Fig. 5a). Therefore, they are tentatively attributed to the intraionic trivalent transitions of Cr3+ placed in a site subjected to a lower crystal field strength than the previous ones. Even though the spectral shape of this Cr3+(4) center is similar to that of Cr3+(3), some differences occur with increasing temperatures. While the latter shows a thermal population when the temperature rises, the Cr3+(4) luminescence arises simultaneously from two split excited states (maxima are separated by 90 meV), and no changes in the peak position were observed with increasing temperature from 70 K to RT. In NIR-I region and on the low energy side of the Cr3+(4) center, we also observe the 4I11/2 → 4I15/2 transition from Er3+, for the ZGeO sample under 457.9 nm excitation.
The presence of multiple Cr3+ optical centers hinder a full deconvolution analysis when temperature-dependent PL measurements are made. Fig. S6a, c and e show the overall PL emission of Cr3+ centers in both samples, between 15 K and RT under 325 nm photon excitation and between 70 K and RT under 532 nm photon excitation. As a general tendency, the overall PL intensity of the RL/NIR-I decreases ca. 72% (Fig. S6b) and 95% (Figures S6d and f) of its original intensity for the ZGeO and ZGeO:Cr samples, respectively.
In the NIR-II spectral region (Fig. 10e), and besides the 4I13/2 → 4I15/2 Er3+ luminescence for the ZGeO sample, the undoped and Cr-doped ZGeO exhibit additional broad bands with the peak position also dependent on the excitation energy and temperature. In particular, at 70 K, by using long wavelength (808 nm) photon excitation, matching the unpolarized absorption bands of Cr4+, a dominant emission with a maximum at 1493 nm/0.83 eV is observed. Hence, it is reasonable to assume that the 1493 nm transition corresponds to the split intraionic 3T2 → 3A2 transition of the Cr4+ in C1 site symmetry (Fig. 6a) with weak crystal field strength, as also reported for the isostructural Zn2SiO4 willemite host.66 In addition, and for the case of the undoped sample, the 1493 nm emission is also detected when 457.9 nm photon excitation (via Cr3+ absorption) is performed. However, if the sample is subjected to 532 nm photon excitation, a shift of the band maximum towards higher energies is noticed, with the emission peak at 1444 nm/0.86 eV. As for the case of Cr3+, such findings suggest that Cr4+ ions could be subjected to slightly different crystalline strengths with sites preferentially excited with different photon energies. As so, we assign the 1493 nm and 1444 nm emission bands to Cr4+(1) and Cr4+(2) centers, respectively, and these are the dominant ones for the ZGeO sample. Furthermore, while the Cr4+(2) emission band keeps its emission maximum independent of temperature, in the case of the Cr4+(1), raising the temperature between 70 K and RT leads to a 30 meV shift of the peak position to high energies, suggesting a thermal population of higher excited states. The temperature-dependent PL spectra of the multiple Cr4+centers of ZGeO sample, as well as the 4I13/2 → 4I15/2 of Er3+ obtained with 532 nm photon excitation, are shown in Figures S7a and c. As for Cr3+ the overlap of the emitting centers (Cr4+ and Er3+) in the measured spectral region hinders a full spectral deconvolution. Therefore, from the overall integrated intensity of all Cr4+ centers, a thermal population occurs between 70 K and 130 K with further nonradiative deexcitation up to RT. As a general tendency, the overall PL intensity decreases ca. 42% of its original intensity for the ZGeO sample (Fig. S7d). In addition, the Er3+ 4I13/2→4I15/2 integrated PL was found to increase by 140% (Fig. S7b), indicating that the emission is thermally populated. For the case of the ZGeO:Cr sample, besides the Cr4+(1) center, exciting the sample with 457.9 nm and 532 nm results, at 70 K, an emission maximum at ca. 1320 nm/0.94 eV with a shoulder at low energy. Such asymmetry suggests that the emission could be due to an overlap between the 1320 nm band (designated as the Cr4+(3) center) and the Cr4+(1) center. Furthermore, raising the temperature between 70 K and RT shows that with 532 nm excitation, the Cr4+(1) is strongly quenched and the emission of the Cr4+(3) center prevails (see Fig. S7e). The overall integrated luminescence of Cr4+centers for the ZGeO:Cr sample center was seen to decrease 85% in the 70 K to RT interval (Fig. S7f). For the 457.9 nm photon excitation, the broadening of the Cr4+(3) center increases, not only due to thermal processes, but also likely due to the contribution of Cr4+(2). Therefore, for the doped sample, three Cr4+ optically active centers are present. Nevertheless, it should be stressed that, as measured by RT absorption (Fig. 5a) and estimated by DFT (Fig. 6a), the Cr4+ ions are in C1 site symmetry when replacing the germanium ions in the ZGeO host. Hence, as happens for Ca2GeO4 and Mg2GeO4,72 the triplet states are expected to unfold under the low site symmetry, giving rise to additional transitions (Fig. S5). As such, we cannot discard that some of the assigned centers could be related to the luminescence of excited states from the same optically active defect. However, and as a main conclusion, our samples clearly evidence that besides the Er3+ contamination in the ZGeO sample, the detected emission in NIR-I and NIR-II spectral regions is due to Cr3+ and Cr4+ that coexist in low site symmetries and weaker crystalline strength on the ZGeO lattice, in good agreement with the theoretical DFT predictions and RT absorption (Fig. 5a).
The presented results are a step forward in the clarification of the nature of the active optical centers in zinc germanate, which evidences a strong dependence on trace contaminants, intrinsic defects, and deliberately introduced chromium dopants, contributing to exploring the host potential as an emerging broad band emitter either in visible and NIR spectral regions.
This journal is © The Royal Society of Chemistry 2025 |