Abstract
Erbium-doped waveguide amplifiers (EDWAs) are vital for photonic integration, yet most are built on z-cut lithium niobate, incompatible with the mainstream x-cut platform. This work presents a combined theoretical and experimental study of polarization-dependent gain in x-cut Er:LNOI. Using Judd–Ofelt theory, we analyze how crystal orientation governs TE-mode coupling to Er3+ ions, predicting stark differences in transition strengths between α- and π-polarizations. Experiments confirm these predictions: at 1,531 nm, the absorption and emission cross sections for α-polarization are 1.8 times larger than for π-polarization. At 1,550 nm, the α-polarization shows a gain coefficient of 3.3 dB/cm versus 2.2 dB/cm for π-polarization. In the small-signal regime, the α-polarized amplifier achieves 32.01 dB signal enhancement with 11.18 dB internal net gain. With 9.1 dBm on-chip input power, it delivers 21.18 mW unsaturated output power under pumping levels exceeding 200 mW. This work demonstrates feasible optical amplification on x-cut LNOI, providing crucial support for large-scale photonic and microwave photonic systems.
1 Introduction
As a key component in modern optical communication systems, erbium-doped fiber amplifiers (EDFAs) have provided essential technological support for the development of global ultra-broadband fiber networks, owing to their high gain, low noise, and excellent compatibility [1]. However, with the rapid advancement of integrated photonics, on-chip active devices are playing an increasingly important role. Among them, on-chip optical amplifiers are a critical element for realizing large-scale integration of active and passive photonic circuits. Building on the success of EDFAs, EDWAs offer effective chip-scale gain to overcome signal attenuation in on-chip optical interconnects [2], making them indispensable for realizing high-performance integrated photonic systems of the future.
Over the past few decades, EDWAs have been extensively investigated in various thin film materials, including silicon nitride [2], [3], [4], aluminum oxide [5], tellurium oxide [6], silicate [7], etc. However, achieving compact EDWAs with high gain remains challenging on these platforms. Recently, LNOI has emerged as a promising platform for multifunctional integrated photonic circuits due to its wide transparent window, excellent electro-optic properties, high nonlinear coefficient, and compatibility with micro–nano fabrication [8], [9], [10], [11]. A wide range of high-performance devices based on LNOI have been demonstrated, such as high-speed electro-optic modulators [12], [13], [14], [15], [16], optical frequency combs [17], [18], [19], efficient wavelength converters [20], [21], and on-chip spectrometers [22], providing advances in large-scale optical interconnects, optical computing, and quantum information processing. Against this backdrop, developing erbium-doped lithium niobate waveguide amplifiers compatible with the LNOI platform would not only enable efficient on-chip signal amplification but also pave the way for the monolithic integration of laser, modulation, and amplification functionalities, which is crucial for future low-power, high-bandwidth-density photonic computing and communication chips.
Several studies have reported EDWAs based on Er:LNOI [23], [24], [25], [26], [27], [28], [29], specifically, Chen et al. has achieved 5.2 dB small signal internal net gain at 1,530 nm [23], Liang et al. has achieved 20 dB small signal internal net gain at 1,532 nm [26], and Cai et al. has reported an output power of 16.65 dBm with 7.65 dB internal net gain at 1,550 nm [29]. These results highlight the potential of this platform for integrating active devices; however, they are all based on z-cut LNOI, whereas x-cut LNOI remains the mainstream choice for most photonic components. In particular, high-speed modulators and nonlinear converters rely on exploiting the strongest electro-optic coefficient (r33) along the x-axis. The inherent differences in crystal orientation between z-cut and x-cut LNOI fundamentally limit seamless integration, creating a pivotal bottleneck for large-scale active photonic systems. Therefore, developing EDWAs base on x-cut LNOI is essential to bridge this technological gap and enable fully integrated photonic circuits.
Here, we focus on x-cut LNOI and systematically investigate the polarization-dependent gain characteristics of erbium-doped lithium niobate waveguide amplifiers. Based on Judd–Ofelt theory, we analyze the variations in Er3+ transition strength and rate-equation parameters for different polarizations and experimentally verify their influence on amplifier performance. This study fills a critical gap in polarization-dependent amplification for x-cut Er:LNOI and provides new insight for cointegrating amplifiers and modulators on a unified LNOI platform, paving the way for high-performance active photonic circuits.
2 Polarization-resolved Judd–Ofelt analysis in x-cut Er:LNOI amplifier
The optical transitions of rare-earth ions in doped crystals are typically analyzed using the Judd–Ofelt (J–O) theory [30], [31]. This semi-empirical model enables the extraction of the J–O intensity parameters (Ω2, Ω4, Ω6) from polarization-resolved absorption spectra, allowing for the calculation of transition strengths between arbitrary energy levels of a given dopant–host system. These parameters are essential for predicting key laser and amplifier performance metrics, such as radiative transition probabilities, radiative lifetimes, and branching ratios.
Lithium niobate (LN) belongs to the trigonal crystal system with space group R3c, and its optical axis typically lies along the z-axis (also referred to as the c-axis). In the undoped structure, both Li+ and Nb5+ ions are coordinated by six oxygen ions, forming distorted octahedral units, as illustrated in the upper panel of Figure 1(a). Upon erbium doping, most Er3+ ions substitute Li+ lattice sites or occupy nearby interstitial positions, while a smaller fraction may incorporate into Nb5+, NbLi4+, and defect cluster positions and so on [32], [33], as shown in the lower panel of Figure 1(a).

Structural and polarization characteristics of x-cut Er3+-doped LN and its relevant energy-level transitions. (a) Crystal structure of undoped LN (left) and a diagonal cut through the hexagonal unit cell of Er3+-doped LN, illustrating the possible substitutional positions of Er3+. (b) π-polarization, light propagates along the y-axis with the electric field oriented parallel to the optical axis (c-axis). (c) α-polarization, light propagates along the z-axis with the electric field oriented perpendicular to the optical axis. (d) Energy model of Er3+ under 1,480 nm pumping.
Due to the anisotropy of LN, its optical properties depend on the relative orientation between the electric field direction and the optical axis. When the electric field is perpendicular to the c-axis, the light propagates as an ordinary ray with refractive index n 0; when it is parallel, the light behaves as an extraordinary ray with refractive index n e [32], [34]. This anisotropy leads to different propagation constants and field distributions under distinct polarizations, which directly affect the coupling efficiency between the external field and the transition dipole of Er3+ ions.
It has been experimentally demonstrated that Er3+:LiNbO3 exhibits strongly polarization-dependent absorption spectra [34], with distinct features for α-polarization (k‖c), σ-polarization (k⊥c, E⊥c), and π-polarization (k⊥c, E‖c).
Within the J–O framework, the electric-dipole line strength (S q,ed) and magnetic-dipole line strength (S q,md) from an initial level a to a final level b under different polarization states are as follows [30], [31]
where e is the electronic charge, Ω
t,q
is the J–O parameters,
The theoretical transition strength from level a to b can be expressed by [31]
where
By measuring the absorption spectra for different polarization states, the experimental transition strengths can be obtained using the expression [31]
where N is the number of dopant ions per unit volume, OD(λ)
i
is the measure optical density under different polarization, defined as
The Ω t,q are then determined by least-squares fitting to the theoretical transition strengths, and the strength of any radiative transition can be calculated.
According to the Ladenburg–Fuchtbauer (L–F) relation [33], the emission cross section (σ em,q ) can be directly expressed as a function of the transition strength
where Φ(λ) is the normalized emission line-shape, and λ is the emission wavelength of interest.
From the above relationship, it is evident that the emission cross section is proportional to the transition strength. Since the J–O parameters Ω t,q are polarization dependent, the electric dipole line strengths differ under different polarizations. In addition, variations in refractive indices lead to changes in the local field correction factors, further modifying the transition strength. As a result, the emission cross section exhibits a clear polarization dependence. According to the McCumber relation [35], the absorption cross section is also polarization dependent.
For x-cut Er:LNOI, the polarization state of the TE mode depends on the propagation direction of the waveguide, as the optical axis lies in the plane of the wafer plane. This is in contrast to z-cut films, where the TE mode always corresponds to a single polarization state (σ-polarization). As illustrated in Figure 1(b) and (c), when the waveguide propagates along the z-axis, the TE mode electric field is perpendicular to the optical axis, corresponding to α-polarization. Conversely, when the waveguide propagates along the y-axis, the TE mode electric field is parallel to the optical axis, corresponding to π-polarization.
Under 1,480 nm pumping, the gain dynamics of Er3+-doped waveguide amplifiers (EDWAs) on lithium niobate on insulator (LNOI) can be described using a steady-state three-level rate equation model [29], as shown in Figure 1(d). Based on the steady-state rate equations, both the net stimulated absorption rate of the pump light and the net stimulated emission rate of the signal light are directly proportional to the corresponding absorption and emission cross sections. Thus, the anisotropy of the J–O parameters directly gives rise to polarization-dependent amplifier gain.
Furthermore, we calculated the transition strength using the J–O parameters reported in Ref. [30] for different polarizations. Values of the host-independent matrix elements as tabulated for Er3+ ions in Ref. [36] were used, and the refractive indices have been obtained from the Sellmeier equation for LiNbO3 reported in the literature [37]. The EDWAs primarily involve the transition between the 4I15/2 and 4I13/2 energy levels under 1,480 nm pumping. The transition strength comprises both electric dipole and magnetic dipole components. Based on Eqs. (1)– (5), the calculation results are shown in Table 1. It should be noted that, due to experimental limitations, we were unable to measure multiple absorption bands to extract J–O parameters in this work, and thus the parameters reported in the literature [30] were used as references. Moreover, since both α and σ polarizations correspond to ordinary rays with refractive index n 0, their J–O parameters can be considered approximately equivalent; therefore, the σ-polarization J–O parameters from the literature were adopted as a proxy for the α-polarization.
Calculated transition strengths for different polarizations. All values are in units of 10−6.
| Transition | α-polarization | π-polarization |
|---|---|---|
| 4I15/2 → 4I13/2 | 1.36 | 1.15 |
From the calculation results, the transition strength of the α-polarization is larger than that of the π-polarization, corresponding to larger absorption and emission cross sections. Therefore, the amplification performance under α polarization is likely to be better than that under π-polarization, which provides theoretical support for the subsequent experimental results.
3 Device design and fabrication
In this work, we designed two polarization-resolved EDWAs on x-cut Er:LNOI. The schematic of EDWAs is shown in Figure 2, which comprise a series of 9 μm-wide ridge waveguides as the gain region and 1 μm-wide ridge waveguides for TE00 mode input/output coupling, interconnected by 1,500 μm-long adiabatic tapers to preserve fundamental mode propagation without mode conversion or mixing. To ensure polarization stability and reduce the overall footprint, a folded waveguide geometry with compact Euler bends was implemented. The Euler bend waveguides have a width of 1 μm, which can simultaneously suppressing higher-order modes and reducing propagation losses. For centimeter-scale amplifiers, the contribution of bending loss can be considered negligible, or approximated as equivalent to the straight-waveguide propagation loss. Figure 2 also shows the simulated mode field profiles at different positions along the waveguide when the input is TE00 mode. The simulation results indicate that the optical field remains predominantly in the fundamental mode throughout the taper and Euler bend regions, with no significant excitation of higher-order modes or polarization conversion.

Schematic of the on-chip erbium-doped waveguide amplifier based on x-cut LNOI. The device features a folded waveguide geometry designed for TE00-mode propagation.
Fabrication process of the EDWA starts from a high quality x-cut Er:LNOI wafer with an erbium doping concentration of 0.5 mol%, which is briefly explained in Figure 3(a). The wafer consists of a 600 nm-thick erbium-doped lithium niobate thin film, a 10 μm-thick SiO2 buffer layer, and a 500 μm-thick silicon substrate. First, electron-beam lithography (EBL) was employed to define the waveguide mask patterns. The mask patterns were then transferred into the Er:LNOI layer using inductively coupled plasma reactive ion etching (ICP-RIE), resulting in ridge waveguides with a 300 nm etching depth. Subsequently, the wafer was annealed at 500 °C to repair lattice damage induced during etching. A 1,000 nm-thick SiO2 cladding layer was then deposited over the waveguides via plasma-enhanced chemical vapor deposition (PECVD). Finally, we cleaved and polished both sides of the chip for edge coupling. Figure 3(b)–(e) shows scanning electron microscope (SEM) images of the EDWA at various locations, including straight waveguides, Euler-bend waveguides, adiabatic tapers, and waveguide sidewalls, all exhibiting smooth surfaces with approximately 65° inclination angles.

Fabrication and structural characterization of the EDWA. (a) Illustration of the fabrication of EDWA base on x-cut LNOI. (b, c, and d) SEM images of the fabricated EDWA at different locations, including the (b) cross sections of the straight waveguides, top views of the (c) adiabatic tapers and (d) Euler bends, (e) the sidewall of the waveguides.
4 Result and discussion
4.1 Absorption loss under different polarizations
In order to accurately estimate the internal net gain of the EDWA, it is essential to properly evaluate the waveguide loss, mainly including the erbium absorption loss and propagation loss caused by waveguide sidewall roughness. Based on the preceding J–O analysis, TE modes propagating in x-cut Er:LNOI waveguides correspond to two distinct polarization states of the Er3+ ions. Due to the anisotropic crystal field, these polarizations exhibit different absorption and emission cross sections, which in turn lead to polarization-dependent absorption losses. The absorption loss
where
We first characterized the coupling and propagation losses of EDWAs at 1,580 nm using the cut-back method, where the absorption of Er3+ ions is negligible [27], [34]. The experimental setup is illustrated in Figure 4(a), mainly consists of a C-band tunable laser source (TSL-550), a polarization controller, and an optical power meter (MPM-210H). This approach was adopted instead of using a ring resonator, since the x-cut Er:LNOI platform is polarization sensitive and a resonator cannot reliably distinguish the losses of α- and π-polarizations. To further minimize the potential influence of residual absorption, the source output power was set to 0 dBm during the measurement.

Characterization of polarization-dependent absorption loss in the EDWA. (a) Experimental setup for measuring the optical total loss. (b) Measured transmission losses at 1,580 nm for two polarizations with varying waveguide lengths, used to extract facet coupling and propagation losses. (c) Small-signal absorption loss spectra across the entire C-band for two EDWA devices with identical geometry but different TE mode polarization states. (d) The loss versus input power at different wavelengths for both polarization states. (e) Measured absorption and emission cross sections of two polarization states under 1,480 nm pumping, with the uncertainty envelopes derived from the cut-back fitting method.
EDWAs with different lengths but the same polarization were measured on the same chip, ensuring that process variations between chips do not affect the results. Figure 4(b) shows the insertion loss versus waveguide length for both α- and π-polarizations. By linear fitting, the facet coupling losses were determined to be 4.88 ± 0.07 dB and 4.59 ± 0.12 dB per facet, and the propagation losses were 0.29 ± 0.01 dB/cm and 0.28 ± 0.02 dB/cm for the α- and π-polarizations, respectively. The small variations indicate that the extracted losses are highly consistent across waveguides of different lengths. Given that the designed device is capable of supporting both pump and signal wavelengths, the subsequent analysis approximates their facet coupling and propagation losses to be identical. For each polarization state, the fitted average values are employed in the calculations.
Then the absorption loss for each polarization can be obtained by subtracting the propagation loss from the on-chip total loss.
To determine the absorption loss under different polarization states, we performed C-band absorption loss measurement on two EDWAs with identical structures but different TE mode polarization orientations – corresponding to distinct erbium dipole alignments (α-polarization and π-polarization). We controlled the above setup via computer to form a rapid wavelength-sweeping measurement system for measuring the output power of the EDWAs across various signal wavelengths.
Figure 4(c) shows the absorption spectra for both polarization states with the source power set to −30 dBm. It is evident that the α-polarized EDWA exhibits significantly higher loss than the π-polarized one. Further comparison with the σ-polarization in z-cut devices shows that the absorption loss of the α-polarization at the peak wavelength of 1,531 nm is also substantially greater than that of both the σ- and π-polarizations.
To further study the power-dependent saturation behavior, we selected several representative wavelengths and measured the absorption loss as a function of input signal power, as presented in Figure 4(d). As the input power increases, the absorption loss gradually decreases, eventually saturating when approximately half of the erbium ions are excited, at which point the net absorption becomes zero. When the input signal power is as small as −30 dBm, the absorption loss no longer increases, approaching the intrinsic Er3+ absorption loss. The reason is that the absorption loss is proportional to the population of Er3+ remaining in the ground state when the absorption cross section is significant. As the signal power increases, more ions are excited, resulting in a reduced ground-state population and hence lower absorption loss. At 1,531 nm, the intrinsic absorption loss is 8.03 dB/cm for the α-polarization and 4.40 dB/cm for the π-polarization. At 1,550 nm, the values are 1.87 dB/cm and 1.22 dB/cm, respectively, indicating that the α-polarization consistently exhibits higher absorption loss than the π-polarization.
According to the McCumber’s theory, the absorption cross section and emission cross section obey the fundamental relation [35]:
where T is the temperature, h is the Planck’s constant, k is Boltzmann’s constant, and ε is the excitation energy related to temperature. In this study, T = 300 K and ɛ = 6,517 cm−1, which are consistent with the parameters reported in Ref. [35].
Therefore, based on the saturated absorption spectra of the two polarizations shown in Figure 4(c), we calculated the actual transition strength, absorption cross section, and emission cross section using Eqs. (6), (8), and (10), respectively.
Table 2 presents a comparison between the experimental transition strengths and the theoretical calculations. It can be observed that the transition intensity for α-polarization is larger than that for π-polarization, which is consistent with the theoretical trend. Meanwhile, the experimental values are generally lower than the theoretical ones, which may be attributed to the fact that the measured absorption spectrum does not fully cover the entire transition range. Nevertheless, the deviations between the experimental and theoretical results for both polarizations fall within the error range reported in the literature [30], indicating the validity and reliability of the results.
Comparison of theoretical and experimental transition strengths under different polarization States. All values are in units of 10−6.
| α-polarization | π-polarization | |||||
|---|---|---|---|---|---|---|
| Transition | f α,th | f α,meas | Δf | f π,th | f π,meas | Δf |
| 4I15/2 → 4I13/2 | 1.36 | 1.30 | 0.06 | 1.15 | 0.89 | 0.26 |
Figure 4(e) shows the absorption and emission cross sections over the entire C-band, together with their corresponding uncertainties. The uncertainties were evaluated based on the facet and propagation loss variances obtained from the cut-back fitting method. Using standard error propagation, it was found that the uncertainties of the absorption cross sections for both polarizations are below 1.9 × 10−22 cm2, while those of the emission cross sections are below 1.17 × 10−21 cm2. Particularly around 1,530 nm and 1,550 nm, these uncertainties are negligible.
Moreover, both the absorption and emission cross sections of the α-polarization are consistently larger than those of the π-polarization across the C-band. At the peak wavelength of 1,531 nm, the absorption and emission cross sections of the α-polarization are approximately 1.8 times those of the π-polarization, highlighting the significant polarization-dependent difference. In particular, at 1,550 nm, the σ abs and σ em of the α-polarization are 0.50 × 10−20 cm2 and 0.73 × 10−20 cm2, respectively, while those of the π-polarization show lower values of 0.33 × 10−20 cm2 and 0.48 × 10−20 cm2. The test results show that the absorption cross section at 1,580 nm is nearly zero, which further supports the validity of treating the loss at 1,580 nm as the propagation loss.
The theoretical prediction is consistent with our experimental results and also agrees with previous findings reported in Er3+-doped LiNbO3 bulk crystals [34], [39].
4.2 Signal enhancement and internal net gain
The measured differences in absorption and emission cross sections suggest that the gain performance of the EDWA should also be polarization-dependent. To verify this, we experimentally characterized the gain for both polarization states. First, we fabricated a series of α-polarized and π-polarized EDWAs with varying lengths on the same chip to evaluate the optimal gain length. The experimental setup is shown in Figure 5. The signal light is from a C-band tunable laser and the pump light is from a 1,480 nm laser diode. The pump light and signal light are combined via an input wavelength division multiplexer (WDM) and then coupled into the EDWA chip through the lensed fiber. The output WDM is used to inject backward pump light and extract the amplified signal light, which is then fed into an optical spectrum analyzer (OSA) for spectral measurement. Both signal light and pump light are adjusted to TE00 mode by polarization controllers (PC) to minimize propagation loss. Notably, under 1,480 nm pumping, the entire EDWA becomes uniformly illuminated by green upconversion fluorescence, visually verifying both the active erbium distribution and effective light confinement within the fabricated waveguides.

Experimental setup to characterize the gain of the EDWA. The digital camera photograph of the excited EDWA chip is shown in the center.
The signal enhancement G se of the EDWA is quantified by measuring the output signal power difference between the pumped and unpumped states, expressed as:
where P on and P off represent the output signal power with and without pumping. This value provides an intuitive representation of the extent of signal enhancement under pumping conditions; however, the internal net gain should be regarded as the primary metric for evaluating actual amplifier performance.
The internal net gain is then obtained by subtracting the total loss from G se, expressed as:
The net gain calculation method employed in this work is widely adopted for the characterization of erbium-doped lithium niobate waveguide amplifiers [23], [24], [25], [26], [27], [28], [29]. Unless otherwise specified, all gain values discussed herein refer to the on-chip internal net gain.
Figure 6(a) describes the internal net gain versus EDWA length for both polarization states under the same sufficient pump power excitation with a fixed on-chip input signal power of −5 dBm (all power values referenced to on-chip levels). The α-polarized and π-polarized configurations achieve nearly identical maximum net gains of 9.87 dB and 10.04 dB at their respective optimal waveguide lengths of 9.71 cm and 11.71 cm. Furthermore, we investigated the relationship between input signal power and gain characteristics. Figure 6(b) shows the internal net gain and output power as a function of input signal power for both polarization states at their respective optimal waveguide lengths. As the input signal power increases, the population inversion decreases, causing the internal net gain to gradually reduce. At first glance, the gain performance of the two polarization states appears similar, showing no significant polarization-dependent behavior. However, it’s worth noting that waveguide lengths for the α-polarization and π-polarization are different. As a result, their gain per unit length and signal enhancement are not directly equivalent. As shown in Figure 6(c) and (d), with an input signal power of approximately −15 dBm, the α-polarized EDWA exhibits 32.01 dB signal enhancement with 3.3 dB/cm unit gain, while the π-polarized EDWA reaches 25.77 dB signal enhancement with 2.2 dB/cm unit gain. According to the rate equation model, the net absorption rate of the pump and the net stimulated emission rate of the signal are both proportional to the absorption and emission cross sections. Therefore, α-polarization can achieve a higher population inversion within a shorter waveguide length, resulting in a larger gain coefficient than π-polarization and consequently a shorter optimum gain length.

On-chip gain performance of the EDWA under different polarization states. (a) Internal net gain of α-polarized and π-polarized EDWAs with varying waveguide lengths under a fixed input signal power of −5 dBm at 1,550 nm. (b) Internal net gain and output signal power versus input signal power at 1,550 nm for both polarization states. (c) and (d) Spectra of the output signal power with and without pump at the maximum gain point for (c) α-polarization and (d) π-polarization. Insets show photos of the device with pump applied.
To further investigate the evolution of signal gain with increasing pump power under different polarizations, we measured the gain evolution as a function of on-chip pump power under different input signal power levels at 1,550 nm and conducted model-based analysis using the experimental data.
As shown in Figure 7(a) and (b), the results show that under small-signal conditions, the α polarization, due to its larger stimulated absorption cross section and higher energy transfer up-conversion coefficient, tends to experience insufficient population inversion at high pump power, where the inversion becomes difficult to sustain and saturation occurs earlier, leading to a rapid gain roll-off. In contrast, under high input signal power, the strong signal significantly depletes the excited-state population, which makes the system more dependent on the pump. In this regime, although the α polarization still has a larger stimulated absorption cross section, it can maintain a relatively stable population inversion at high pump powers, and thus does not exhibit the same gain degradation observed in the small-signal regime.

Amplifier performance under different pump power levels. Measured (scatters) and simulated (curves) internal net gain with respect to the on-chip pump power at three input signal powers for (a) α-polarization and (b) π-polarization; along with the measured on-chip output signal power for (c) α-polarization and (d) π-polarization.
For the π-polarization, due to its relatively smaller stimulated absorption cross section and lower energy transfer up-conversion coefficient, it can maintain a more stable population inversion under small-signal conditions, resulting in smoother gain growth at low to moderate pump powers. However, when the input signal power is high, the signal significantly depletes the inversion, making the system more reliant on strong pumping to replenish the excited-state population. Although the pump power increases, the π-polarization exhibits a relatively smaller absorption cross section for the pump light, and the longer propagation length further reduces the effective pumping efficiency. As a result, even under high pump powers, it is difficult to maintain sufficient population inversion, leading to gain saturation and eventual decline.
Figure 7(c) and (d) shows the on-chip output power as a function of on-chip pump power. At an input signal power of 9.1 dBm, the α-polarized device achieves a maximum on-chip output power of 21.18 mW, and the π-polarized counterpart reaches 17.82 mW. However, based on the observed trend of the output power curves, it can be expected that under increased pump power, α-polarization still has the potential to further enhance output power, exhibiting a higher saturation gain limit. This higher saturation gain limit and pump demand are attributed to a faster inversion depletion rate in the α-polarization, resulting from its larger absolute absorption cross section coupled with stronger excited-state absorption and upconversion.
5 Conclusions
5.1 Polarization-dependent gain performance and mechanism
The measured absorption and emission cross sections exhibit a pronounced disparity between α- and π-polarizations. At the peak wavelength of 1,531 nm, the cross sections for α-polarization are 1.8 times greater than those for π-polarization. More notably, at the standard telecommunication wavelength of 1,550 nm, the absorption and emission cross sections were quantified as 0.50 × 10−20 cm2 and 0.73 × 10−20 cm2 for α-polarization, compared to 0.33 × 10−20 cm2 and 0.48 × 10−20 cm2 for π-polarization, respectively.
This substantial difference in cross sections directly leads to divergent amplifier gain performance. The higher cross sections of the α-polarization result in a higher gain coefficient of 3.3 dB/cm, allowing it to achieve optimal amplification in a shorter device length of 9.71 cm. In contrast, the π-polarization, with its lower gain coefficient of 2.2 dB/cm, requires a longer interaction length of 11.71 cm to attain comparable gain levels. Consequently, under high input signal power conditions, the α-polarized amplifier delivers a superior maximum unsaturated output power of 21.18 mW, outperforming the 17.82 mW achieved by the π-polarized device.
These performance discrepancies originate fundamentally from the crystalline anisotropy of the x-cut LNOI substrate, which strongly influences the transition strengths of the Er3+ ions. Judd–Ofelt theoretical calculations confirm that the transition strength for α-polarization is significantly higher than that for π-polarization. This directly predicts the larger cross sections observed experimentally and provides a rigorous physical explanation for the enhanced gain and output performance of the α-polarization across the C-band.
5.2 Comparison with conventional z-cut platform
To evaluate the implications of our platform choice, we compare the performance of our x-cut amplifier with the well-established z-cut platform in Table 3. The comparison reveals a clear trade-off: while z-cut EDWAs generally yield higher absolute net gain and output power [24], [25], [26], [28], [29], their performance is largely isotropic. In contrast, the x-cut platform provides a decisive new degree of freedom – polarization-dependent gain – as demonstrated in this work. More importantly, it offers superior electro-optic (EO) and nonlinear optic (NLO) compatibility due to facile access to the dominant r33 and d33 coefficients, which is a fundamental requirement for future densely integrated photonic circuits.
Performance comparison between conventional z-cut and this work on x-cut Er:LNOI waveguide amplifiersa.
| Ref. | Orientation | Polarizationb | Device length (cm) | Pump wavelength (nm) | Pump power (mW) | Signal wavelength (nm) | Input signal power (dBm) | Internal net gain (dB)c | Output power (dBm) | EO/NLO compatibility | Implication/trade-off |
|---|---|---|---|---|---|---|---|---|---|---|---|
| [24] | Z-cut | TE (σ-pol) | 3.6 | 980 | 40 | 1,530 | −47.07 | 18 | −29.07 | Poor (relies on r13) |
Z-cut platforms generally achieve higher absolute gain |
| 26 | 1,550 | −38.38 | 7.5 | −30.88 | |||||||
| [25] | Z-cut | TE (σ-pol) | 2.58 | 1,484 | 17.31 | 1,531.6 | −50 | 16 | −34 | Poor (relies on r13) |
X-cut offers polarization-sensitive gain, a key feature for advanced control |
| 18.54 | 1,550 | −50 | 2.66 | −47.34 | |||||||
| [26] | Z-cut | TE (σ-pol) | 10 | 976 | 63 | 1,532 | −22 | 20 | −2 | Poor (relies on r13) |
α-pol requires a shorter device length for peak performance |
| [28] | Z-cut | TE (σ-pol) | 7 | 1,486 | 210 | 1,550 | 4.5 | 16 | 20.5 | Poor (relies on r13) |
π-pol requires significantly lower pump power |
| 63 | 1,531.5 | −33 | 27.6 | −5.4 | |||||||
| [29] | Z-cut | TE (σ-pol) | 10 | 1,484 | N/A | 1,531 | −15 | 22.26 | 7.26 | Poor (relies on r13) |
Z-cut enables higher output power levels |
| 120 | 1,550 | 9 | 7.65 | 16.65 | |||||||
| This work | X-cut | TE (α-pol) | 9.71 | 1,480 | 65 | 1,550 | −15 | 11.18 | −3.82 | Excellent (full r33 acess) |
X-cut is superior for monolithic integration with modulators & wavelength converters |
| 206 | 1,550 | 9.1 | 4.16 | 13.26 | |||||||
| This work | X-cut | TE (π-pol) | 11.71 | 1,480 | 65 | 1,550 | −15 | 11.30 | −3.7 | Excellent (full r33 acess) |
|
| 110 | 1,550 | 9.1 | 3.41 | 12.51 |
-
aAll results are approximate values obtained from the reported literature. bThe α-, σ-, and π-polarizations refer to the Er3+ transition orientations determined by the electric-field and propagation directions relative to the optical axis. cAll gain values correspond to the total on-chip internal net gain, excluding coupling losses.
5.3 Advancements beyond prior work on x-cut LNOI
Having contextualized the x-cut platform against the z-cut standard, we now highlight the specific advancements of our work within the emerging field of x-cut Er:LNOI amplifiers.
As summarized in Table 4, pioneering studies by Luo et al. [40] and Xue et al. [27] successfully demonstrated the feasibility of achieving optical gain on this platform. However, their analyses were not designed to resolve the fundamental polarization-dependent characteristics inherent to the x-cut orientation.
Comparison with prior work on x-cut Er:LNOI waveguide amplifiers.
| Ref. | Gain reported | Polarization resolved | Key achievement | Advancement of this work |
|---|---|---|---|---|
| [23] | Yes | No | First demonstration of an EDWA on x-cut LNOI, proving feasibility | Provides the first comprehensive polarization-resolved analysis of gain, cross sections, and loss |
| [27] | Yes | No | First amplifier fabricated on um-thick x-cut LNOI, demonstrating higher gain | First quantification of polarization-dependent gain coefficients (3.3 vs 2.2 dB/cm) |
| This work | Yes | Yes | Systematic characterization of polarization-dependent spectral and amplification properties | Supplies key data and design rules for developing polarization-aware integrated functional devices |
Our work provides the first comprehensive, polarization-resolved analysis of transition strength, gain, cross sections, and saturation properties on this platform. This dataset provides the indispensable foundation for designing sophisticated, polarization-aware functional devices on this platform.
5.4 Conclusion and outlook
In summary, this work systematically investigated the polarization-dependent gain characteristics of EDWAs on the x-cut LNOI platform. Guided by J–O theory, we analyzed the variations in Er3+ transition strength and rate-equation parameters for different polarizations, revealing the underlying mechanism of how crystal orientation and polarization state affect device performance. Experimental validation confirmed significant differences in absorption/emission cross sections and gain coefficients between α- and π-polarizations, leading to distinct trends in optimal gain length, gain saturation, and output performance. A maximum unsaturated output power of 21.18 mW is achieved under 1,480 nm pumping in a 9.71 cm-long α-polarized EDWA, indicating the potential for even higher output power with increased pump power. This study provides the first comprehensive insight into polarization-selective amplification in x-cut Er:LNOI and establishes both theoretical and experimental foundations for designing efficient, fully integrated EDWAs, paving the way for large-scale photonic integration on a unified photonic platform.
Funding source: National Natural Science Foundation of China
Award Identifier / Grant number: 62405308
-
Research funding: This research was funded by National Nature Science Fund of China under Grants 62405308.
-
Author contribution: JH conducted the device design, experimental measurements, data analysis, and manuscript writing. RL carried out numerical simulations and participated in the discussion of the results. SW provided theoretical guidance and contributed to manuscript writing and editing. QJ and ZX contributed to the fabrication of the sample. JY contributed to the early stages of the work. JL and JL supervised the project and provided overall guidance throughout the research. All authors have accepted responsibility for the entire content of this manuscript and consented to its submission to the journal, reviewed all the results, and approved the final version of the manuscript.
-
Conflict of interest: Authors state no conflict of interest.
-
Data availability: The datasets generated and analyzed during the current study are available from the corresponding authors upon reasonable request.
References
[1] P. M. Becker, A. A. Olsson, and J. R. Simpson, Erbium-Doped Fiber Amplifiers: Fundamentals and Technology, San Diego, Academic Press, 1999.10.1016/B978-012084590-3/50007-7Suche in Google Scholar
[2] Y. Liu et al.., “A photonic integrated circuit–based erbium-doped amplifier,” Science, vol. 376, no. 6599, pp. 1309–1313, 2022, https://doi.org/10.1126/science.abo2631.Suche in Google Scholar PubMed
[3] J. Mu, M. Dijkstra, J. Korterik, H. Offerhaus, and S. M. García-Blanco, “High-gain waveguide amplifiers in Si3N4 technology via double-layer monolithic integration,” Photonics Res., vol. 8, no. 10, pp. 1634–1641, 2020, https://doi.org/10.1364/prj.401055.Suche in Google Scholar
[4] J. Rönn et al.., “Ultra-high on-chip optical gain in erbium-based hybrid slot waveguides,” Nat. Commun., vol. 10, no. 1, p. 432, 2019, https://doi.org/10.1038/s41467-019-08369-w.Suche in Google Scholar PubMed PubMed Central
[5] D. B. Bonneville, C. E. Osornio-Martinez, M. Dijkstra, and S. M. García-Blanco, “High on-chip gain spiral Al2O3:Er3+ waveguide amplifiers,” Opt. Express, vol. 32, no. 9, pp. 15527–15536, 2024, https://doi.org/10.1364/oe.516705.Suche in Google Scholar PubMed
[6] H. C. Frankis et al.., “Erbium-doped TeO2-coated Si3N4 waveguide amplifiers with 5 dB net gain,” Photonics Res., vol. 8, no. 2, pp. 127–134, 2020, https://doi.org/10.1364/prj.8.000127.Suche in Google Scholar
[7] S. A. Vázquez-Córdova et al.., “Erbium-doped spiral amplifiers with 20 dB of net gain on silicon,” Opt. Express, vol. 22, no. 21, pp. 25993–26004, 2014, https://doi.org/10.1364/oe.22.025993.Suche in Google Scholar
[8] J. Lin, F. Bo, Y. Cheng, and J. Xu, “Advances in on-chip photonic devices based on lithium niobate on insulator,” Photonics Res., vol. 8, no. 12, pp. 1910–1936, 2020, https://doi.org/10.1364/prj.395305.Suche in Google Scholar
[9] D. Zhu et al.., “Integrated photonics on thin-film lithium niobate,” Adv. Opt. Photonics, vol. 13, no. 2, pp. 242–352, 2021, https://doi.org/10.1364/aop.411024.Suche in Google Scholar
[10] A. Boes et al.., “Lithium niobate photonics: unlocking the electromagnetic spectrum,” Science, vol. 379, no. 6627, p. eabj4396, 2023, https://doi.org/10.1126/science.abj4396.Suche in Google Scholar PubMed
[11] M. G. Vazimali and S. Fathpour, “Applications of thin-film lithium niobate in nonlinear integrated photonics,” Adv. Photonics, vol. 4, no. 3, p. 034001, 2022, https://doi.org/10.1117/1.ap.4.3.034001.Suche in Google Scholar
[12] C. Wang et al.., “Integrated lithium niobate electro-optic modulators operating at CMOS-compatible voltages,” Nature, vol. 562, no. 7725, pp. 101–104, 2018, https://doi.org/10.1038/s41586-018-0551-y.Suche in Google Scholar PubMed
[13] M. Xu et al.., “High-performance coherent optical modulators based on thin-film lithium niobate platform,” Nat. Commun., vol. 11, no. 1, p. 3911, 2020, https://doi.org/10.1038/s41467-020-17806-0.Suche in Google Scholar PubMed PubMed Central
[14] P. Kharel, C. Reimer, K. Luke, L. He, and M. Zhang, “Breaking voltage-bandwidth limits in integrated lithium niobate modulators using micro-structured electrodes: erratum,” Optica, vol. 8, no. 9, p. 1, 2021, https://doi.org/10.1364/optica.440484.Suche in Google Scholar
[15] Y. Xue et al.., “Breaking the bandwidth limit of a high-quality-factor ring modulator based on thin-film lithium niobate,” Optica, vol. 9, no. 10, pp. 1131–1137, 2022, https://doi.org/10.1364/optica.470596.Suche in Google Scholar
[16] J. Liu et al.., “Lithium niobate thin film electro-optic modulator,” Nanophotonics, vol. 13, no. 8, pp. 1503–1508, 2024, https://doi.org/10.1515/nanoph-2023-0865.Suche in Google Scholar PubMed PubMed Central
[17] M. Zhang et al.., “Broadband electro-optic frequency comb generation in a lithium niobate microring resonator,” Nature, vol. 568, no. 7752, pp. 373–377, 2019, https://doi.org/10.1038/s41586-019-1008-7.Suche in Google Scholar PubMed
[18] Y. Hu, C. Reimer, A. Shams-Ansari, M. Zhang, and M. Loncar, “Realization of high-dimensional frequency crystals in electro-optic microcombs,” Optica, vol. 7, no. 9, pp. 1189–1194, 2020, https://doi.org/10.1364/optica.395114.Suche in Google Scholar
[19] T. H. Wu et al.., “Visible-to-ultraviolet frequency comb generation in lithium niobate nanophotonic waveguides,” Nat. Photonics, vol. 18, no. 3, 2024, https://doi.org/10.1038/s41566-023-01364-0.Suche in Google Scholar
[20] J.-Y. Chen, Z.-H. Ma, Y. M. Sua, Z. Li, C. Tang, and Y.-P. Huang, “Ultra-efficient frequency conversion in quasi-phase-matched lithium niobate microrings,” Optica, vol. 6, no. 9, pp. 1244–1245, 2019, https://doi.org/10.1364/optica.6.001244.Suche in Google Scholar
[21] C. Wang et al.., “Ultrahigh-efficiency wavelength conversion in nanophotonic periodically poled lithium niobate waveguides,” Optica, vol. 5, no. 11, pp. 1438–1441, 2018, https://doi.org/10.1364/optica.5.001438.Suche in Google Scholar
[22] D. Pohl et al.., “An integrated broadband spectrometer on thin-film lithium niobate,” Nat. Photonics, vol. 14, no. 1, pp. 24–29, 2020, https://doi.org/10.1038/s41566-019-0529-9.Suche in Google Scholar
[23] Z. X. Chen et al.., “Efficient erbium-doped thin-film lithium niobate waveguide amplifiers,” Opt. Lett., vol. 46, no. 5, pp. 1161–1164, 2021, https://doi.org/10.1364/ol.420250.Suche in Google Scholar
[24] J. X. Zhou et al.., “On-Chip integrated waveguide amplifiers on erbium-doped thin-film lithium niobate on insulator,” Laser Photonics Rev., vol. 15, no. 8, p. 2100030, 2021, https://doi.org/10.1002/lpor.202100030.Suche in Google Scholar
[25] M. L. Cai et al.., “Erbium-doped lithium niobate thin film waveguide amplifier with 16 dB internal net gain,” IEEE J. Sel. Top. Quant. Electron., vol. 28, no. 3, p. 8200608, 2022, https://doi.org/10.1109/jstqe.2021.3137192.Suche in Google Scholar
[26] Y. T. Liang et al.., “A high-gain cladded waveguide amplifier on erbium doped thin-film lithium niobate fabricated using photolithography assisted chemo-mechanical etching,” Nanophotonics, vol. 11, no. 5, pp. 1033–1040, 2022, https://doi.org/10.1515/nanoph-2021-0737.Suche in Google Scholar PubMed PubMed Central
[27] X. Xue et al.., “Integrated erbium-doped waveguide amplifier on lithium niobate on insulator,” Opt. Mater. Express, vol. 14, no. 8, p. 10, 2024, https://doi.org/10.1364/ome.532439.Suche in Google Scholar
[28] R. Bao et al.., “An erbium-doped waveguide amplifier on thin film lithium niobate with an output power exceeding 100 Mw,” Laser Photonics Rev., no. 1, p. 19, 2025.10.1002/lpor.202400765Suche in Google Scholar
[29] M. Cai et al.., “Gain dynamics in integrated waveguide amplifier based on erbium-doped thin-film lithium niobate,” ACS Photonics, vol. 11, no. 11, pp. 4923–4932, 2024, https://doi.org/10.1021/acsphotonics.4c01433.Suche in Google Scholar
[30] L. Núñez, G. Lifante, and F. Cussó, “Polarization effects on the line-strength calculations of Er3+-doped LiNbO3,” Appl. Phys. B, vol. 62, no. 5, pp. 485–491, 1996, https://doi.org/10.1007/bf01081048.Suche in Google Scholar
[31] J. Amin, B. Dussardier, T. Schweizer, and M. Hempstead, “Spectroscopic analysis of Er3+ transitions in lithium niobate,” J. Lumin., vol. 69, no. 1, pp. 17–26, 1996, https://doi.org/10.1016/0022-2313(96)00063-4.Suche in Google Scholar
[32] D. M. Gill, L. McCaughan, and J. C. Wright, “Spectroscopic site determinations in erbium-doped lithium niobate,” Phys. Rev. B, vol. 53, no. 5, pp. 2334–2344, 1996, https://doi.org/10.1103/physrevb.53.2334.Suche in Google Scholar PubMed
[33] V. Dierolf and M. Koerdt, “Combined excitation-emission spectroscopy of Er3+ ions in stoichiometric LiNbO3: The site selectivity of direct and up conversion excitation processes,” Phys. Rev. B, vol. 61, no. 12, pp. 8043–8052, 2000, https://doi.org/10.1103/physrevb.61.8043.Suche in Google Scholar
[34] C. H. Huang, L. McCaughan, and D. M. Gill, “Evalution of absorption and emission cross sections of Er-Doped LINBO3 for application to integrated optic amplifiers,” J. Lightwave Technol., vol. 12, no. 5, pp. 803–809, 1994, https://doi.org/10.1109/50.293972.Suche in Google Scholar
[35] D. L. Zhang, W. B. Sun, W. H. Wong, and Y. B. Pun, “Emission and absorption cross sections of Er3+:LiNbO3 crystal: composition effect,” Opt. Mater. Express, vol. 5, no. 9, p. 1920, 2015, https://doi.org/10.1364/ome.5.001920.Suche in Google Scholar
[36] W. T. Carnall, P. R. Fields, and K. Rajnak, “Electronic energy levels in the trivalent lanthanide aquo ions. I. Pr3+, Nd3+, Pm3+, Sm3+, Dy3+, Ho3+, Er3+, and Tm3+,” J. Chem. Phys., vol. 49, no. 10, pp. 4424–4442, 1968, https://doi.org/10.1063/1.1669893.Suche in Google Scholar
[37] M. V. Hobden and J. Warner, “The temperature dependence of the refractive indices of pure lithium niobate,” Phys. Lett., vol. 22, no. 3, pp. 243–244, 1966, https://doi.org/10.1016/0031-9163(66)90591-9.Suche in Google Scholar
[38] J. D. B. Bradley, Al2O3:Er3+ as a Gain Platform for Integrated Optics, Enschede, University of Twente, 2009.Suche in Google Scholar
[39] M. Dinand and W. Sohler, “Theoretical modeling of optical amplification in Er-doped Ti:LiNbO3 waveguides,” IEEE J. Quantum Electron., vol. 30, no. 5, pp. 1267–1276, 1994, https://doi.org/10.1109/3.303692.Suche in Google Scholar
[40] Q. Luo et al.., “On-chip erbium-doped lithium niobate waveguide amplifiers [Invited],” Chin. Opt. Lett., vol. 19, no. 6, p. 06008, 2021, https://doi.org/10.3788/col202119.060008.Suche in Google Scholar
© 2025 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Artikel in diesem Heft
- Frontmatter
- Reviews
- Light-driven micro/nanobots
- Tunable BIC metamaterials with Dirac semimetals
- Large-scale silicon photonics switches for AI/ML interconnections based on a 300-mm CMOS pilot line
- Perspective
- Density-functional tight binding meets Maxwell: unraveling the mysteries of (strong) light–matter coupling efficiently
- Letters
- Broadband on-chip spectral sensing via directly integrated narrowband plasmonic filters for computational multispectral imaging
- Sub-100 nm manipulation of blue light over a large field of view using Si nanolens array
- Tunable bound states in the continuum through hybridization of 1D and 2D metasurfaces
- Integrated array of coupled exciton–polariton condensates
- Disentangling the absorption lineshape of methylene blue for nanocavity strong coupling
- Research Articles
- Demonstration of multiple-wavelength-band photonic integrated circuits using a silicon and silicon nitride 2.5D integration method
- Inverse-designed gyrotropic scatterers for non-reciprocal analog computing
- Highly sensitive broadband photodetector based on PtSe2 photothermal effect and fiber harmonic Vernier effect
- Online training and pruning of multi-wavelength photonic neural networks
- Robust transport of high-speed data in a topological valley Hall insulator
- Engineering super- and sub-radiant hybrid plasmons in a tunable graphene frame-heptamer metasurface
- Near-unity fueling light into a single plasmonic nanocavity
- Polarization-dependent gain characterization in x-cut LNOI erbium-doped waveguide amplifiers
- Intramodal stimulated Brillouin scattering in suspended AlN waveguides
- Single-shot Stokes polarimetry of plasmon-coupled single-molecule fluorescence
- Metastructure-enabled scalable multiple mode-order converters: conceptual design and demonstration in direct-access add/drop multiplexing systems
- High-sensitivity U-shaped biosensor for rabbit IgG detection based on PDA/AuNPs/PDA sandwich structure
- Deep-learning-based polarization-dependent switching metasurface in dual-band for optical communication
- A nonlocal metasurface for optical edge detection in the far-field
- Coexistence of weak and strong coupling in a photonic molecule through dissipative coupling to a quantum dot
- Mitigate the variation of energy band gap with electric field induced by quantum confinement Stark effect via a gradient quantum system for frequency-stable laser diodes
- Orthogonal canalized polaritons via coupling graphene plasmon and phonon polaritons of hBN metasurface
- Dual-polarization electromagnetic window simultaneously with extreme in-band angle-stability and out-of-band RCS reduction empowered by flip-coding metasurface
- Record-level, exceptionally broadband borophene-based absorber with near-perfect absorption: design and comparison with a graphene-based counterpart
- Generalized non-Hermitian Hamiltonian for guided resonances in photonic crystal slabs
- A 10× continuously zoomable metalens system with super-wide field of view and near-diffraction–limited resolution
- Continuously tunable broadband adiabatic coupler for programmable photonic processors
- Diffraction order-engineered polarization-dependent silicon nano-antennas metagrating for compact subtissue Mueller microscopy
- Lithography-free subwavelength metacoatings for high thermal radiation background camouflage empowered by deep neural network
- Multicolor nanoring arrays with uniform and decoupled scattering for augmented reality displays
- Permittivity-asymmetric qBIC metasurfaces for refractive index sensing
- Theory of dynamical superradiance in organic materials
- Second-harmonic generation in NbOI2-integrated silicon nitride microdisk resonators
- A comprehensive study of plasmonic mode hybridization in gold nanoparticle-over-mirror (NPoM) arrays
- Foundry-enabled wafer-scale characterization and modeling of silicon photonic DWDM links
- Rough Fabry–Perot cavity: a vastly multi-scale numerical problem
- Classification of quantum-spin-hall topological phase in 2D photonic continuous media using electromagnetic parameters
- Light-guided spectral sculpting in chiral azobenzene-doped cholesteric liquid crystals for reconfigurable narrowband unpolarized light sources
- Modelling Purcell enhancement of metasurfaces supporting quasi-bound states in the continuum
- Ultranarrow polaritonic cavities formed by one-dimensional junctions of two-dimensional in-plane heterostructures
- Bridging the scalability gap in van der Waals light guiding with high refractive index MoTe2
- Ultrafast optical modulation of vibrational strong coupling in ReCl(CO)3(2,2-bipyridine)
- Chirality-driven all-optical image differentiation
- Wafer-scale CMOS foundry silicon-on-insulator devices for integrated temporal pulse compression
- Monolithic temperature-insensitive high-Q Ta2O5 microdisk resonator
- Nanogap-enhanced terahertz suppression of superconductivity
- Large-gap cascaded Moiré metasurfaces enabling switchable bright-field and phase-contrast imaging compatible with coherent and incoherent light
- Synergistic enhancement of magneto-optical response in cobalt-based metasurfaces via plasmonic, lattice, and cavity modes
- Scalable unitary computing using time-parallelized photonic lattices
- Diffusion model-based inverse design of photonic crystals for customized refraction
- Wafer-scale integration of photonic integrated circuits and atomic vapor cells
- Optical see-through augmented reality via inverse-designed waveguide couplers
- One-dimensional dielectric grating structure for plasmonic coupling and routing
- MCP-enabled LLM for meta-optics inverse design: leveraging differentiable solver without LLM expertise
- Broadband variable beamsplitter made of a subwavelength-thick metamaterial
- Scaling-dependent tunability of spin-driven photocurrents in magnetic metamaterials
- AI-based analysis algorithm incorporating nanoscale structural variations and measurement-angle misalignment in spectroscopic ellipsometry
Artikel in diesem Heft
- Frontmatter
- Reviews
- Light-driven micro/nanobots
- Tunable BIC metamaterials with Dirac semimetals
- Large-scale silicon photonics switches for AI/ML interconnections based on a 300-mm CMOS pilot line
- Perspective
- Density-functional tight binding meets Maxwell: unraveling the mysteries of (strong) light–matter coupling efficiently
- Letters
- Broadband on-chip spectral sensing via directly integrated narrowband plasmonic filters for computational multispectral imaging
- Sub-100 nm manipulation of blue light over a large field of view using Si nanolens array
- Tunable bound states in the continuum through hybridization of 1D and 2D metasurfaces
- Integrated array of coupled exciton–polariton condensates
- Disentangling the absorption lineshape of methylene blue for nanocavity strong coupling
- Research Articles
- Demonstration of multiple-wavelength-band photonic integrated circuits using a silicon and silicon nitride 2.5D integration method
- Inverse-designed gyrotropic scatterers for non-reciprocal analog computing
- Highly sensitive broadband photodetector based on PtSe2 photothermal effect and fiber harmonic Vernier effect
- Online training and pruning of multi-wavelength photonic neural networks
- Robust transport of high-speed data in a topological valley Hall insulator
- Engineering super- and sub-radiant hybrid plasmons in a tunable graphene frame-heptamer metasurface
- Near-unity fueling light into a single plasmonic nanocavity
- Polarization-dependent gain characterization in x-cut LNOI erbium-doped waveguide amplifiers
- Intramodal stimulated Brillouin scattering in suspended AlN waveguides
- Single-shot Stokes polarimetry of plasmon-coupled single-molecule fluorescence
- Metastructure-enabled scalable multiple mode-order converters: conceptual design and demonstration in direct-access add/drop multiplexing systems
- High-sensitivity U-shaped biosensor for rabbit IgG detection based on PDA/AuNPs/PDA sandwich structure
- Deep-learning-based polarization-dependent switching metasurface in dual-band for optical communication
- A nonlocal metasurface for optical edge detection in the far-field
- Coexistence of weak and strong coupling in a photonic molecule through dissipative coupling to a quantum dot
- Mitigate the variation of energy band gap with electric field induced by quantum confinement Stark effect via a gradient quantum system for frequency-stable laser diodes
- Orthogonal canalized polaritons via coupling graphene plasmon and phonon polaritons of hBN metasurface
- Dual-polarization electromagnetic window simultaneously with extreme in-band angle-stability and out-of-band RCS reduction empowered by flip-coding metasurface
- Record-level, exceptionally broadband borophene-based absorber with near-perfect absorption: design and comparison with a graphene-based counterpart
- Generalized non-Hermitian Hamiltonian for guided resonances in photonic crystal slabs
- A 10× continuously zoomable metalens system with super-wide field of view and near-diffraction–limited resolution
- Continuously tunable broadband adiabatic coupler for programmable photonic processors
- Diffraction order-engineered polarization-dependent silicon nano-antennas metagrating for compact subtissue Mueller microscopy
- Lithography-free subwavelength metacoatings for high thermal radiation background camouflage empowered by deep neural network
- Multicolor nanoring arrays with uniform and decoupled scattering for augmented reality displays
- Permittivity-asymmetric qBIC metasurfaces for refractive index sensing
- Theory of dynamical superradiance in organic materials
- Second-harmonic generation in NbOI2-integrated silicon nitride microdisk resonators
- A comprehensive study of plasmonic mode hybridization in gold nanoparticle-over-mirror (NPoM) arrays
- Foundry-enabled wafer-scale characterization and modeling of silicon photonic DWDM links
- Rough Fabry–Perot cavity: a vastly multi-scale numerical problem
- Classification of quantum-spin-hall topological phase in 2D photonic continuous media using electromagnetic parameters
- Light-guided spectral sculpting in chiral azobenzene-doped cholesteric liquid crystals for reconfigurable narrowband unpolarized light sources
- Modelling Purcell enhancement of metasurfaces supporting quasi-bound states in the continuum
- Ultranarrow polaritonic cavities formed by one-dimensional junctions of two-dimensional in-plane heterostructures
- Bridging the scalability gap in van der Waals light guiding with high refractive index MoTe2
- Ultrafast optical modulation of vibrational strong coupling in ReCl(CO)3(2,2-bipyridine)
- Chirality-driven all-optical image differentiation
- Wafer-scale CMOS foundry silicon-on-insulator devices for integrated temporal pulse compression
- Monolithic temperature-insensitive high-Q Ta2O5 microdisk resonator
- Nanogap-enhanced terahertz suppression of superconductivity
- Large-gap cascaded Moiré metasurfaces enabling switchable bright-field and phase-contrast imaging compatible with coherent and incoherent light
- Synergistic enhancement of magneto-optical response in cobalt-based metasurfaces via plasmonic, lattice, and cavity modes
- Scalable unitary computing using time-parallelized photonic lattices
- Diffusion model-based inverse design of photonic crystals for customized refraction
- Wafer-scale integration of photonic integrated circuits and atomic vapor cells
- Optical see-through augmented reality via inverse-designed waveguide couplers
- One-dimensional dielectric grating structure for plasmonic coupling and routing
- MCP-enabled LLM for meta-optics inverse design: leveraging differentiable solver without LLM expertise
- Broadband variable beamsplitter made of a subwavelength-thick metamaterial
- Scaling-dependent tunability of spin-driven photocurrents in magnetic metamaterials
- AI-based analysis algorithm incorporating nanoscale structural variations and measurement-angle misalignment in spectroscopic ellipsometry