Erbium-doped TiO2 materials are promising candidates for advancing quantum technologies, necessitating a thorough understanding of their electronic and crystal structures to tailor their properties and enhance coherence times. This study explored epitaxial erbium-doped rutile TiO2 films deposited on r-sapphire substrates using molecular beam epitaxy. Photoluminescence excitation spectroscopy demonstrated decreasing photoluminescence lifetimes with erbium doping, indicating limited optical coherence times. Lattice distortions associated with Er3+ were probed by x-ray absorption spectroscopy, indicating that erbium primarily occupies Ti4+ sites and influences oxygen vacancies. Significant lattice distortions in the higher-order shells and apparent full coordination around erbium suggest that additional defects are likely prevalent in these regions. These findings indicate that defects contribute to limited coherence times by introducing alternative decay pathways, leading to shorter photoluminescence lifetimes.
I. INTRODUCTION
The relentless growth in data storage and the continuous trend of device miniaturization has driven technological progress to approach the quantum limit.1,2 Quantum information science (QIS), grounded in the principles of quantum mechanics, emerges as a groundbreaking solution for efficient data handling, secure information exchange, and long-distance communication while ensuring data integrity.3,4 At the heart of quantum technologies lies the manipulation of qubits, the fundamental units of quantum information.4,5 Quantum coherence plays a crucial role in realizing quantum technologies,6 which enables qubits to maintain the superposition of quantum states. However, the challenge lies in preserving coherence, which is limited by the inherent susceptibility to environmental influences. Therefore, ensuring the prolonged persistence of quantum states in superposition (i.e., long coherence time) is essential for successfully implementing quantum information technologies.5,6
Solid-state spin qubits show significant promise among the candidates for QIS applications, owing to their compatibility with conventional microelectronics.5,6 Nevertheless, the properties of solid-state spin qubits are linked to their crystal host, necessitating material engineering to enhance the quantum emitter’s properties and mitigate decoherence sources such as charge carriers or phonons. Among potential candidates, rare-earth ions (REIs) in solid-state systems stand out.7 In particular, the trivalent erbium ion (Er3+) is noteworthy for its optical 4f–4f transition within telecom-C band, aligning with the low-loss transmission in optical fibers, making it an ideal quantum emitter for quantum network systems.8–13 Selecting the optimal crystal host for REIs in QIS involves considering factors such as low nuclear and electron spin concentrations, nonpolar symmetry, and a large bandgap.6,13,14 These factors are instrumental in not only enhancing optical transitions but also in mitigating local effects in the vicinity of REIs, thereby addressing coherence issues effectively. Previous studies identified TiO2 as a promising host for Er3+, with the rutile and anatase phases being particularly stable.15,16 Recent research on Er-doped TiO2 films and bulk crystals has demonstrated narrow optical inhomogeneous linewidths, underscoring the importance of buffer and capping layers in enhancing optical properties.15,17 Despite these advancements, further understanding is needed on how Er3+ dopants influence the TiO2 host lattice’s electronic and crystal structure.
This investigation focuses on epitaxial erbium-doped rutile TiO2 thin films grown on r-plane sapphire substrates, aiming to explore how erbium incorporation affects the material’s optical properties. We employed spectroscopy and diffraction techniques to correlate structural modifications with variations in excited state lifetimes. The selection of sapphire as the substrate was deliberate, as it enables the formation of high-quality single-crystal rutile TiO2. X-ray diffraction (XRD) confirmed the persistence of single crystallinity despite doping. Variations on the Er3+ photoluminescence lifetimes, which set the upper bounds for optical coherence times, were assessed using photoluminescence excitation (PLE) spectroscopy revealing a downward trend with increasing doping. In addition, information on local coordination, lattice distortion, and chemical environment near erbium sites was probed using x-ray absorption spectroscopy (XAS) and extended x-ray absorption fine structure (EXAFS). The analyses indicate that doping increases oxygen vacancy defect density within TiO2. Furthermore, they suggest that erbium ions are fully coordinated, indicating redistribution of oxygen vacancies away from the Er sites. This increase in additional defect density is likely to limit coherence by introducing alternative decay pathways upon Er3+ excitation.
II. METHODS
A. Samples Growth
Epitaxial Er:TiO2 thin films with thicknesses of 65 nm were deposited using a Riber C21 DZ Cluster molecular beam epitaxy (MBE) system. Titanium tetraisopropoxide (TTIP) from Sigma-Aldrich with a purity of 99.999% (trace metal basis) was the titanium precursor as in the TTIP-based MBE of TiO2 detailed elsewhere.18 Growths were carried out on an r-plane (102) sapphire substrate at 640 °C temperature, oxygen partial pressure of 7 × 10−6 Torr, and a TTIP beam-equivalent flux of 3.5 × 10−6. The growth rate was 32.5 nm per hour. In situ erbium (Er) doping (20, 200, and 500 ppm) using metallic Er was carried out by a high-temperature (860–990 °C) effusion cell.
B. Structural and optical characterization
Single crystal x-ray diffraction characterization was carried out at the 33BM beamline of the Advanced Photon Source (APS) at photon energy of 20 keV. PLE spectroscopy was performed in a custom confocal microscopy setup designed for telecom C-band measurement of thin films, with samples mounted in a cryostat at 3.5 K (s50 Cryostation, Montana Instruments). A tunable 1.5 μm laser illuminated the samples for 1.5 ms using pulses shaped by acousto-optic modulators. Emission from the samples subsequent to excitation was then collected by a Quantum Opus superconducting nanowire single photon detector (SNSPD) for 3 ms collection intervals. Sweeping the excitation laser wavelength resulted in a PLE spectrum and excited state lifetime measurements at each wavelength for each measured sample. Additional information on this experimental setup is given in detail elsewhere.12 XAS measurements were conducted at O K-edge, Ti L-edge, and Er M5-edges at the 29ID beamline at the Advanced Photon Source (APS). The beam was directed along surface normal, and the spectra were averaged over the acquired horizontal and vertical polarization to ensure orientation-independent assessment of the local environment. The EXAFS measurements were performed at the 20-BM beamline of the APS. Data collection was conducted in grazing incidence geometry, with an angle of incidence of a few degrees and with linear x-ray polarization in the plane of the film. The sample was rotated about the film surface normal to average in-plane polarizations. The temperature was maintained at 300 K throughout the entire experiment. Since the [101] direction is along the surface normal, separate FEFF8 simulations were carried out with linear polarization along orthogonal in-plane [−101] and [010] directions to obtain theoretical phase shifts and scattering amplitudes.19
III. RESULTS AND DISCUSSION
A. Characterization of epitaxial Er:TiO2 films by single crystal x-ray diffraction
Figure 1 shows the single crystal x-ray diffraction spectra, confirming the direction normal to the r-plane surface through the specular (0, K, 2K) scan. The film shows a main peak of film orientation on the r-plane surface consistent with previous literature.20,21 Positions of the peaks are roughly equivalent with the expected bulk values indicating that the films are relaxed. The lattice parameter approximates that of bulk TiO2. As was noted in a previous work,20 the matching of the symmetry between the and planes leads to a twinned structure with alternating domains on 10’s of nm length scales. These domains likely lead to a relaxation to the bulk lattice values due to the high-density of domain walls. In addition, all films exhibit a minority (301) orientation, which may result from the high density of twin domains. No systematic behavior with doping was observed, suggesting that the minority (301) orientation is unrelated to doping.
B. Impact of erbium concentration on optical properties
The impact of increasing Er3+ dopant density on the rutile TiO2 host lattice was investigated by optical characterization of the thin film samples performed via time-resolved PLE spectroscopy. Figure 2(a) shows the resulting absorption spectra near 1.5 μm resonance for the series of single crystal erbium-doped rutile TiO2 films. The ground and excited electronic states of Er3+ in rutile fully split into eight (Z1−8) and seven (Y1−7) Kramers’ doublets, respectively [Fig. 2(b)],15 due to the D2h point symmetry for Er3+ occupying Ti4+ sites in rutile TiO2. The growing background toward lower wavelengths stems from the dopant-driven introduction of additional defects and disorder into the lattice, resulting in the available crystal field transitions broadening significantly.22 At high doping levels (such as the 500 ppm sample in this work), the inhomogeneous linewidths are broad enough that individual peaks become more difficult to resolve.
We focus our discussion on the peak highlighted by a vertical dashed line (∼1520 nm), which corresponds to the technologically prescient22 Er3+ Z1 → Y1 transition in rutile-phase TiO2.15 The narrowest observed inhomogeneous linewidth in this set of thin films is ∼50 GHz, found in the sample doped with 20 ppm erbium. The inhomogeneous linewidth increases with higher dopant concentrations, with the broadening of the linewidth attributable to electric fields generated by charged defects near the Er3+ sites, including negatively charged point defects resulting from Er3+ substituting Ti4+ sites (i.e., the negative charge upon replacement of Ti4+ to Er3+) and positively charged oxygen vacancy point defects.17,23–26
The lifetime of the excited state sets an upper limit for the coherence time, and shorter lifetimes may also be an indicator of lower quantum efficiency. Therefore, identifying dopant-driven variation in the excited state lifetime is critical in assessing the material’s quality as a host for quantum processes, given the intentional introduction of dopants, such as Er3+, as necessary to facilitate these processes. Understanding the nuances of local electronic and crystal structures is pivotal for comprehending the alterations in the inherent optical properties of Er-doped rutile TiO2 with varying erbium concentrations.
The photoluminescence lifetimes at the highlighted peak (Table I) exhibit a dopant-driven decreasing trend. This shortening of lifetime is consistent with prior observations in CeO2 with increasing erbium doping, where the decrease is attributed to the introduction of defects. While the exact nature of these defects remains uncertain, they likely introduce non-radiative or quenching pathways that shorten the photoluminescence lifetimes.22 If the defects were strictly localized to individual Er sites, their interaction would be limited and neither significant linewidth broadening nor a pronounced reduction in lifetime would be expected. However, the impact of these defects is significant, and is emphasized by the decrease in quantum efficiency with higher doping (also presented in Table I), estimated based on the longest observed excited state lifetime of Y1 → Z1 for Er-doped rutile TiO2 in the literature, of 5.25 ± 0.03 ms.15 The increased linewidth and shorter photoluminescence lifetimes at higher dopant concentrations suggest that the defects surrounding the Er3+ sites are not confined to the immediate vicinity but instead form more extended distributions. These extended defects likely modify the local crystal field around the Er ions, thereby influencing the observed optical properties.
C. Influence of erbium on the electronic structure of TiO2 host lattice
To further elucidate these modifications, we employed XAS to investigate erbium-induced changes in the local electronic structure within the crystal lattice. Starting with oxygen spectroscopy, the O K-edge spectra shown in Fig. 3(a) are a result of electronic transitions from O(1s) states to unoccupied O(2p) states. The spectra were acquired simultaneously in both total electron yield (TEY) mode, using the sample photocurrent method, and fluorescence yield (FY) mode, using an energy-discriminating silicon drift diode detector (Vortex) coupled with a multi-channel analyzer (MCA). The TEY and FY spectra are represented by the solid and dashed lines, respectively. The primary distinction between TEY and FY acquisition modes lies in their depth sensitivity, with TEY being more surface-sensitive and FY more bulk-sensitive. TiO2 consists of a central Ti4+ ion surrounded by six O2− ions arranged as a distorted TiO6 octahedron due to the tetragonal symmetry of the crystal structure. This distortion causes the crystal field splitting of the Ti 3d orbitals into lower-energy t2g states and higher-energy eg states.27–29 Geometrically, the eg orbitals from Ti point toward the O 2p orbitals, while the t2g lobes point between the O atoms leading to strong hybridization between O(2p) and Ti(3d) orbitals. The significant hybridization leads to two distinguishable pre-peaks at ∼530 and 533 eV labeled as t2g and eg, respectively. Moreover, three distinct features at higher energies, with peaks around 539, 542, and 545 eV, denoted as C, D, and E, respectively, correspond to delocalized O(2p)–Ti(4s) and O(2p)–Ti(3d) hybridized states and are a fingerprint of the rutile phase in TiO2.30
Two notable trends emerge from comparing the TEY and FY spectra. First, the intensity ratio of the t2g and eg pre-peaks varies with increasing erbium concentration in both TEY and FY spectra, suggesting that these variations are not confined to the surface and exist in the bulk of the film as well. Second, the FY spectra, which provide greater depth sensitivity than TEY, display broader peaks between 538 and 545 eV. This broadening is attributed to contributions from the sapphire substrate.31 Due to the strong hybridization between O(2p) and Ti(3d) orbitals, the t2g and eg pre-peaks in the O K-edge spectra are expected to be highly sensitive to changes in the electronic structure. Thus, the fitting analysis shown in Fig. 3(b) focused on the two pre-peaks in the TEY spectra to gain insights into the electronic structure of the host lattice. Gaussian functions were fitted to the spectra, revealing two distinct features labeled as A* and B*, adjacent to the main t2g and eg peaks, respectively. These adjacent features are ∼1 eV blueshifted from the main peaks and can be attributed to modifications in the electronic structure of the rutile lattice related to oxygen vacancies.32
Notably, the A* and B* features are present even in the undoped sample, indicating non-stoichiometric TiO2. Their area increases with erbium doping, as presented in Table II. Native point defects, such as neutral or charged oxygen vacancies, are common in wide-bandgap metal oxides such as TiO2.33,34 The increase in A* and B* areas with doping suggests that erbium introduces new oxygen vacancies into the lattice. Each Er ion is expected to create only half an oxygen vacancy to balance the charge (i.e., one vacancy balances the charge of two Er3+ ions). Despite this, the substantial increase in A* and B* peak areas implies that defects around different Er sites likely overlap or interact. This overlap leads to significant changes in the XAS spectra, even at low Er concentrations, suggesting that the defects are likely not isolated but form a more extensive network within the material. Moreover, the erbium-driven increase in A* and B* areas is accompanied by a decrease in the t2g and eg peak areas, reflecting disruptions in the local electronic structure. The creation of new oxygen vacancies disrupts the strong hybridization between O(2p) and Ti(3d) orbitals, leading to reduced overlap and thus smaller t2g and eg peak areas.
Er concentration (ppm) . | Peak area (a.u.) . | |||
---|---|---|---|---|
. | t2g . | A* . | eg . | B* . |
0 | 1.22 (14) | 0.47 (9) | 1.13 (8) | 0.24 (7) |
20 | 0.86 (14) | 0.76 (9) | 0.98 (8) | 0.33 (7) |
200 | 0.77 (14) | 0.73 (9) | 0.84 (8) | 0.40 (7) |
Er concentration (ppm) . | Peak area (a.u.) . | |||
---|---|---|---|---|
. | t2g . | A* . | eg . | B* . |
0 | 1.22 (14) | 0.47 (9) | 1.13 (8) | 0.24 (7) |
20 | 0.86 (14) | 0.76 (9) | 0.98 (8) | 0.33 (7) |
200 | 0.77 (14) | 0.73 (9) | 0.84 (8) | 0.40 (7) |
Similar to the O K-edge, the Ti L–edge XAS spectra are sensitive to changes in electronic structure due to the hybridization of O(2p) and Ti(3d) orbitals. Therefore, the Ti L–edge was probed using TEY mode to investigate possible dopant-induced modifications in the electronic structure (Fig. 4). The fine structure in the L3 and L2 edges results from the breaking of degenerate electronic orbitals by the lattice crystal field, leading to the splitting of 3d states into threefold t2g and twofold degenerate eg states. Further distortion from the lattice environment introduces additional splitting, which is pronounced at the L3-edge. The first two weak pre-edge peaks are attributed to the combined effects of particle–hole coupling and crystal field splitting. These features exhibit low intensity due to partial screening of the core hole by surrounding electrons, which is influenced by the localized electronic environment and transition metal oxidation state, as discussed in previous studies.35,36 The intense, narrow peak around 458 eV corresponds to the t2g state, while the eg states split into two peaks due to the distortion in the TiO6 octahedra. This distortion leads to an inversion of the relative intensities of the eg peaks in the XAS spectra for different TiO2 polymorphs. As a result, the relative intensity ratio of the eg peaks serves as a distinct fingerprint for identifying TiO2 polymorphs, and in this case, it is consistent with the rutile phase.28
Expected consequences of doping include charge compensation through the formation of additional defects, such as oxygen vacancies and Ti3+, as well as distortions of the host lattice.26,37,38 Prior investigations on Cr-doped rutile TiO2 highlighted the critical role of oxygen point defects in stabilizing Cr3+ ions, particularly through the formation of complex defects involving two Cr atoms and one oxygen vacancy.38 This stabilization is associated with distortions in the crystal lattice due to the elongation of Ti–O bonds, which affect the Ti eg orbitals. These orbitals are highly sensitive to changes in the local environment because they interact directly with the 2p orbitals of the surrounding O atoms. As the local symmetry and crystal field around the Ti ions change, the energy levels of the eg orbitals shift accordingly. In our study, the low concentration of Er3+ dopant—on the order of hundreds of parts per million (ppm)—makes it challenging to observe significant lattice distortions or prominent features associated with eg orbital splitting or Ti3+-related signals. However, comparisons between undoped and doped samples reveal an overall decrease in spectral intensity with Er doping, suggesting that the dopant induces changes in the local electronic environment around the Ti ions. These changes may arise from slight distortions in the Ti–O coordination environment, which, in turn, alter the crystal field and electronic structure of the Ti sites.
To gain deeper insight into how the dopant affects the formation of additional defects within the host lattice, erbium spectroscopy was performed. The XAS spectrum at the Er M5-edge (Fig. 5) exhibits three distinct peaks corresponding to 3d → 4f transitions. These transitions, characterized by changes in the total angular momentum quantum number (ΔJ = 0, ±1), are indicative of erbium in the trivalent state.39
D. Structural distortions upon erbium incorporation into TiO2
The local environment around the Er3+ sites was further examined using EXAFS at the Er L-edge and Ti K-edge, as shown in Fig. 6. A comparison of the experimental data with FEFF8 simulations is shown in Figs. 6(a) and 6(d), where the Fourier-transformed (FT) XAFS data for the Ti K-edge and Er L3-edge, respectively, are represented by the solid lines.
These comparisons revealed significant differences, particularly in the FT Er L3-edge spectra, where a marked shift toward higher distances in the first shell was observed, along with a notable suppression in the higher-R structure. These observations suggest alterations in the local structure around erbium, such as bond length variations or changes in coordination numbers at greater distances. The experimental data, acquired at 300 K, were scaled by a factor of 2 in these two plots to better match with simulations conducted at 0 K.
Figures 6(e) and 6(b) show FT XAFS data and fits for Er L3-edge and Ti K-edge, respectively. Each polarization contribution was equally weighted (0.5) in the fits to account for in-plane polarization and sample spin, employed during the data acquisition to reduce Bragg peak artifacts in the experiment. The analysis involved simultaneous fitting of data weighted by different k-weights to minimize correlations between coordination number and the Debye–Waller factor (σ2). The small nominal split of 0.04 Å in first shell Ti/Er–O distances cannot be resolved with the limited k-range of the XAFS data (kmax ∼ 10 Å−1) hence the fitted bond length correction was forced to be proportional to the nominal bond lengths. Similarly, coordination numbers of split distances were fitted while maintaining their ratio fixed to nominal values. The FT Er L3-edge spectrum [Fig. 6(e)] revealed erbium coordinated to oxygen for the first shell (Er–O), with an expanded bond length of 2.20 Å (Table III). This distance is significantly larger than the corresponding Ti–O(1) bond length of 1.92 Å observed in the Ti K-edge analysis, indicating a local expansion in the lattice around the erbium dopant. In addition, the presence of Ti neighbors at a distance of 3.09 Å in the second shell suggests that Er3+ is occupying Ti4+ lattice sites and is consistent with previous ESR-based investigations on Er-doped rutile TiO215.
Probed atom . | Bond . | Bond length (Å) . | Debye waller factor (Å2) . | Coordination number . | |||
---|---|---|---|---|---|---|---|
. | . | 0 ppm . | 500 ppm . | 0 ppm . | 500 ppm . | 0 ppm . | 500 ppm . |
Ti K-edge | Ti–O(1) | 1.934 (5)a | 1.92 (1) | 0.0055 (5) | 0.0060 (1) | 4.1 (2) | 4.1 (2) |
Ti–O(2) | 1.978 (5)b | 1.96 (1) | 0.0055 (5) | 0.0060 (1) | 2.1 (1) | 2.1 (1) | |
Ti–Ti(1) | 3.02 (1)c | 3.00 (3) | 0.0040 (1) | 0.0030 (1) | 2d | 2d | |
Ti–Ti(2) | 3.61 (1)e | 3.62 (2) | 0.0062 (9) | 0.0010 (4) | 8d | 8d | |
Er L3-edge | Er–O(1) | ⋯ | 2.20 (1) | ⋯ | 0.0056 (7) | ⋯ | 4.3 (2) |
Er–O(2) | ⋯ | 2.25 (1) | ⋯ | 0.0056 (7) | ⋯ | 2.1 (1) | |
Er–Ti(1) | ⋯ | 3.09 (3) | ⋯ | 0.0140 (5) | ⋯ | 2d |
Probed atom . | Bond . | Bond length (Å) . | Debye waller factor (Å2) . | Coordination number . | |||
---|---|---|---|---|---|---|---|
. | . | 0 ppm . | 500 ppm . | 0 ppm . | 500 ppm . | 0 ppm . | 500 ppm . |
Ti K-edge | Ti–O(1) | 1.934 (5)a | 1.92 (1) | 0.0055 (5) | 0.0060 (1) | 4.1 (2) | 4.1 (2) |
Ti–O(2) | 1.978 (5)b | 1.96 (1) | 0.0055 (5) | 0.0060 (1) | 2.1 (1) | 2.1 (1) | |
Ti–Ti(1) | 3.02 (1)c | 3.00 (3) | 0.0040 (1) | 0.0030 (1) | 2d | 2d | |
Ti–Ti(2) | 3.61 (1)e | 3.62 (2) | 0.0062 (9) | 0.0010 (4) | 8d | 8d | |
Er L3-edge | Er–O(1) | ⋯ | 2.20 (1) | ⋯ | 0.0056 (7) | ⋯ | 4.3 (2) |
Er–O(2) | ⋯ | 2.25 (1) | ⋯ | 0.0056 (7) | ⋯ | 2.1 (1) | |
Er–Ti(1) | ⋯ | 3.09 (3) | ⋯ | 0.0140 (5) | ⋯ | 2d |
1944 Å.
1988 Å.
2959 Å.
Fixed.
357 Å.
The significant local expansion observed around Er aligns with the difference in ionic radii between Er3+ and Ti4+. The theoretical difference in ionic radii is 0.29 Å (0.89–0.605 Å40), which closely matches the experimental local expansion of 0.28 Å, as indicated by the difference between the Er–O(1) bond length (2.20 Å) and the Ti–O(1) bond length (1.92 Å) presented in Table III. This close agreement between theoretical and experimental values, combined with nominal (N = 2) Er–Ti coordination in the second shell at a bond distance of 3.09 Å, supports the hypothesis that Er3+ is substituting at Ti4+ sites, albeit with significant lattice distortion especially at higher order shells. Interestingly, the amplitude of Ti K-edge XAFS is somewhat reduced in the doped sample, despite the rather low doping level of 500 ppm, indicating that structural distortions around Er ions are extended, affecting a larger % of Ti ions. These results can also be correlated with the overall intensity decreasing in Ti L-edge XAS spectra with doping shown in Fig. 4, as erbium ions appear to cause lattice distortions. The Ti K-edge analysis [Fig. 6(b)] for the undoped sample (blue circles) yielded fitted distances closely resembling the nominal TiO2 structure. Small differences were observed for the doped sample (red circles), although signatures of increased disorder are present particularly at the longer Ti–Ti correlations (Table III). The same amplitude reduction factor () of 0.93 was used for both the Ti K-edge and Er L3-edge fits to ensure consistency across the analyses.
The substitution of Ti4+ by Er3+ ions raises important questions about the mechanisms of charge compensation that stabilize the dopant within the host structure. The increased area of the A* and B* features in the O K-edge spectra, which are associated with oxygen vacancies, suggests that these vacancies play a key role in compensating for the charge imbalance introduced by Er3+ ions, thereby maintaining charge neutrality.41,42 EXAFS analysis indicates that the first coordination shell of both Er3+ and Ti4+ remains fully occupied by oxygen, as reflected by coordination numbers close to 6 for both species. In particular, the coordination number of Er–O is 6.4 (3) (Table III), while the coordination number for Ti–O is 6.2 (3), supporting full occupation in the first shell. However, the error bars of around 5% in the coordination number allow for the presence of oxygen vacancies in the first Ti–O shell. Thus, while the first shell appears fully occupied within experimental uncertainty, the possibility of oxygen vacancies beyond the first Er–O shell in the lattice cannot be ruled out.
The σ2 factors indicate increased disorder beyond the first shell, particularly around the Er3+ ions, with a value of 0.0140 (5) Å2 for the Er–Ti(1) distance. In addition, major differences between the data and simulations in the higher shells [Fig. 6(d)] point to erbium-driven structural changes beyond the first coordination shell, which may include oxygen vacancies located farther from the Er3+ sites. We were unable to model the Er local structure beyond the Er–Ti(1) distance. The long-range distortions in the local environment around Er3+ ions, extending to at least 4 Å, suggest that these defects are not isolated but cause widespread changes in the lattice. Even at 500 ppm doping levels, these extensive distortions may overlap, affecting a significant amount of Ti atoms, as suggested by the Ti K- and L-edge probes.
Finally, revisiting the PLE discussion, the observed decrease in the photoluminescence lifetime of Er3+ can be attributed to the dopant-induced generation of oxygen vacancies. These point defects act as non-radiative recombination centers, enabling energy transfer between Er3+ ions and oxygen vacancy defects, which leads to a quenching effect. Consequently, the photoluminescence lifetime decreases due to enhanced non-radiative recombination pathways. Since the photoluminescence lifetime imposes an upper limit on coherence time in quantum systems, the observed reduction in photoluminescence lifetimes indicates a corresponding decrease in coherence times. Other potential sources of non-radiative quenching, such as domain-associated defects or crystal field distortions, may also contribute to this effect and require further investigation to fully elucidate their roles.
IV. CONCLUSION
In conclusion, our investigation into epitaxial Er-doped rutile TiO2 thin films reveals significant effects of erbium doping on both structural and electronic properties. XAS analysis at the O K-edge indicates that erbium doping leads to the formation of oxygen vacancies, which alter the electronic structure of the host lattice. EXAFS analysis for the Er L3-edge spectra indicates that Er3+ ions substitute for Ti4+ ions in the TiO2 lattice. This substitution results in notable lattice distortions due to the larger ionic radius of Er3+ compared to Ti4+. Coordination number analysis reveals consistent sixfold coordination around erbium, suggesting that neighboring oxygen ions stabilize the local charge environment. However, while our findings indicate significant distortions in higher-order shells, which may be associated with oxygen vacancies displaced from the first shell, further comparative studies are necessary to definitively determine the exact substitution sites. The broadening of the f–f excitation observed in PLE suggests a distribution of crystal fields, indicating structural disorder extending beyond the first coordination shell. This disorder impacts the crystal field sensed by the 4f electrons. The lattice distortions and additional point defects are expected to enhance non-radiative energy transfer, which correlates with the reduced photoluminescence lifetimes observed by PLE. This underscores the complex interplay between dopant-induced structural modifications and optical properties, emphasizing the importance of detailed local electronic and crystal structure investigations for advancing quantum information processing technologies.
ACKNOWLEDGMENTS
This work was supported by Q-NEXT, a U.S. Department of Energy Office of Science National Quantum Information Science Research Centers under Award No. DE-FOA-0002253. The use of the Advanced Photon Source, Argonne National Laboratory was supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, under Contract No. DE-AC02-06CH11357.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
J. B. Martins: Conceptualization (equal); Data curation (lead); Formal analysis (equal); Investigation (equal); Methodology (equal); Validation (equal); Visualization (lead); Writing – original draft (lead); Writing – review & editing (equal). G. Grant: Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). D. Haskel: Data curation (equal); Formal analysis (equal); Methodology (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). G. E. Sterbinsky: Data curation (equal); Investigation (supporting); Methodology (equal); Validation (equal); Visualization (equal); Writing – review & editing (equal). I. Masiulionis: Data curation (supporting); Validation (supporting); Visualization (supporting). K. E. Sautter: Conceptualization (equal); Investigation (supporting); Methodology (equal); Visualization (equal). E. Karapetrova: Data curation (equal); Methodology (equal). S. Guha: Funding acquisition (equal); Visualization (equal); Writing – review & editing (equal). J. W. Freeland: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available within this article.