Optica

Optica

Published by Optica Publishing Group

Online ISSN: 2334-2536

Disciplines: Optics

Journal websiteAuthor guidelines

Top-read articles

118 reads in the past 30 days

Effect of coherence on an optical system. ( ${{\rm a}_1}$ )–( ${{\rm a}_3}$ ) Intensity patterns of double-slit interference for coherent, partially coherent, and incoherent light, respectively. ( ${{\rm b}_1}$ )–( ${{\rm b}_3}$ ) Intensity profiles at the output plane corresponding to ( ${{\rm a}_1}$ )–( ${{\rm a}_3}$ ), respectively.
Conceptual sketch of the experimental setup. (a) Illustration of the PC-DONN setup employing spatial light modulators (SLMs). See the text for the details of the setup. (b) The hologram loaded onto ${{\rm SLM}_1}$ to generate the handwritten digits. (c) The input light features the pattern of handwritten digits. (d) The recognition result of differential measurement (the red and blue circles represent the positive and negative detectors, respectively).
Comparisons of traditional coherent (a)–(c) and our partially coherent (d)–(f) DONN. ( ${{\rm a}_1}$ )–( ${{\rm a}_4}$ ) Experimental outputs of the traditional coherent DONN when the coherence lengths of testing light are $l = \infty ,1,0.2,0.05\;{\rm mm}$ , respectively. ( ${{\rm b}_1}$ )–( ${{\rm b}_4}$ ) Corresponding numerical results. ( ${{\rm c}_1}$ )–( ${{\rm c}_4}$ ) Corresponding recognition probabilities for experimental (Exp) and numerical (Num) results. (d)–(f) Respectively correspond to (a)–(c) with the exception that the DONN is a PC-DONN trained at $l = 0.2\;{\rm mm}$ . In ( ${{\rm c}_1}$ )–( ${{\rm c}_4}$ ) and ( ${{\rm f}_1}$ )–( ${{\rm f}_4}$ ), the digits corresponding to the recognition results are highlighted in red. And the reduced visibility in the experimental outputs of the DONN arises due to the different total number of phase screens used between experimental (M = 50) and numerical (M = 1000) setups. (Scale bar: 500 µm).
Images depict the output light from the first layer of coherent DONN and PC-DONN, corresponding to various input lights with different coherence lengths as specified in Table 5. ( ${a_1}$ )–( ${a_5}$ ) Represent the output for a coherence length of $l = 1$ mm. ( ${b_1}$ )–( ${b_5}$ ) Show outputs for coherence lengths of $l = 0.05\,\,\rm mm$ . ( ${c_1}$ )–( ${c_5}$ ) and ( ${d_1}$ )–( ${d_5}$ ) Show outputs of PC-DONN for coherence lengths of $l = 1\,\,\rm mm$ and $l = 0.05\,\,\rm mm$ , respectively.
Partially coherent diffractive optical neural network

December 2024

·

243 Reads

Qi Jia

·

·

Yanxia Zhang

·

[...]

·

Download

Aims and scope


Optica is dedicated to the rapid dissemination of the highest impact results in all areas of optics and photonics. This online-only, Open Access journal publishes original peer-reviewed research articles, letters, memoranda, and mini-reviews that appeal to a broad audience. A highly selective journal, Optica is a venue for authors to publish their most exciting work, be it theoretical or experimental, fundamental or applied. The Journal also provides a transparent peer review option.

Recent articles


Photonic skin for wearable sensing using photonic integration
  • Article

January 2025

Hongqiang Li

·

Xiaolin Li

·

Yueting Yang

·

[...]

·

Enbang Li




Side-coupled 2D optomechanical crystal (OMC) cavity. (a) Helium-ion microscope image of a representative device with insets indicating salient features from left to right: unitcells of the 2D snowflake lattice, central fish-bone waveguide, optical coupling waveguide, and the optical waveguide mirror. The orange inset on the right shows the supercell of the geometry used to simulate optical and acoustic bandstructures. (b) Simulated optical bandstructure of the supercell. The solid red (blue) band has energy predominantly in the cavity (coupling waveguide). Transverse electric field profiles of these optical modes at the X-point are shown on the right. Dashed lines in the bandstructure indicate other guided modes. (c) FEM simulations of the acoustic (left; total displacement) and optical (right; transverse electric field) modes of the 2D-OMC cavity with acoustic resonance frequency, ${\Omega _{\rm m}}/2\pi = 10.3\;{\rm GHz} $ , and optical resonance wavelength, $\lambda = 1550\;{\rm nm} $ , respectively.
Characterization of optical-absorption-induced hot bath. (a) Schematic showing interactions of the acoustic resonator with various baths considered in our heating model. (b) Schematic of measurement setup for time-resolved measurements of the hot bath using single-photon counting on the optical sideband generated by thermal motion of the acoustic resonator. (c) Measurement of the transient thermal occupation of the acoustic resonator, $n_{\rm m}^ \star$ , in response to a rectangular optical pulse on resonance with the optical cavity (pulse duration ${\tau _{\rm d}} = 50\;{\unicode{x00B5} \rm s}$ , repetition rate, $R = 1\;{\rm kHz}$ , and peak intracavity photon occupation, ${n_{\rm c}} = 385$ ). The black line represents an exponential fit to the observed data with the characteristic rate, ${\gamma _{\rm p}} + {\gamma _0}$ , and steady-state thermal occupation, ${n_{\rm m}}$ . Here ${\gamma _0}$ is the intrinsic damping rate of the acoustic resonator and ${\gamma _{\rm p}}$ is the coupling rate to the optical-absorption-induced hot bath. (d) Thermal occupation of the hot bath, ${n_{\rm p}}$ estimated from measurements of ${n_{\rm m}}$ performed at varying optical power, shown on the $x$ -axis in units of peak intra-cavity photon occupation, ${n_{\rm c}}$ . For comparison, ${n_{\text{p}}}$ curves for 1D-OMC [3], butt-coupled 2D-OMC [11] are shown. (e) Variation of ${\gamma _{\rm p}}/2\pi$ with ${n_{\rm c}}$ . The data point marked with an arrow in (d) and (e) corresponds to the data in (c) for ${n_{\rm c}} = 385$ .
Phonon-to-photon transduction under continuous-wave excitation. (a) Schematic of measurement setup showing single-photon counting of up-converted photons at the optical resonance frequency with the OMC pumped continuously on the red-detuned sideband ( $\Delta = - {\Omega _{\rm m}}$ ) of the optical resonance. Measured thermal phonon occupancy, ${n_{{\rm th}}}$ , with varying optical power, shown on the bottom $x$ -axis in units of intra-cavity photon occupation, ${n_{\rm c}}$ , and on the top-axis in units of optomechanical transduction rate, ${\gamma _{{\rm OM}}}$ . Results are shown on separate charts for (b) device I, and (c) device II. Filled circles are data points whereas the solid line indicated with $\beta = 0\;{{\unicode{x00B5} \rm w}^{- 1}}$ is the modeled ${n_{{\rm th}}}$ dependence using Eq. (2). For comparison, model curves are shown for a butt-coupled 2D-OMC ( $\beta = 15\; {{\unicode{x00B5} \rm W}^{- 1}}$ ) [11], and 1D-OMC [3]. Dashed lines indicate the ${n_{\rm c}}$ value for optomechanical transduction efficiency ${\eta _{{\rm OM}}} = 50\%$ and 90%. For ${n_{\rm c}} = 1$ , on-chip input powers for devices I and II are ${P_{{\rm in}}} = 0.20\;{\unicode{x00B5} \rm W}$ and 0.28 µW, respectively. (d) Estimated heating rate of the acoustic resonator ${\gamma _{\rm p}}{n_{\rm p}}/2\pi$ as a function of ${n_{\rm c}}$ under $\Delta = - {\Omega _{\rm m}}$ for different devices plotted for their measured value of $\beta$ .
Phonon-to-photon transduction under pulsed laser excitation. Internal added noise, $\bar n$ (red data points), as a function of the peak intra-cavity pump photon number, ${n_{\rm c}}$ . All noise measurements are performed on device II with rectangular optical pump pulses with a pulse width of 500 ns at a repetition rate of 250 Hz. For comparison, $\bar n$ is shown for butt-coupling design [11], and 1D-OMC [3]. Dashed lines indicate the ${n_{\rm c}}$ value for transduction efficiency ${\eta _{{\rm OM}}} = 10\%$ , 50%, and 90% for device II. For ${n_{\rm c}} = 1$ , on-chip input power is ${P_{{\rm in}}} = 0.28\;{\unicode{x00B5} \rm W}$ .
High-efficiency low-noise optomechanical crystal photon-phonon transducers
  • Article
  • Full-text available

January 2025

·

20 Reads

·

3 Citations

Optomechanical crystals (OMCs) enable coherent interactions between optical photons and microwave acoustic phonons, and represent a platform for implementing quantum transduction between microwave and optical signals. Optical-absorption-induced thermal noise at cryogenic (millikelvin) temperatures is one of the primary limitations of performance for OMC-based quantum transducers. Here, we address this challenge with a two-dimensional silicon OMC resonator that is side-coupled to a mechanically detached optical waveguide, realizing a six-fold reduction in the heating rate of the acoustic resonator compared to prior state-of-the-art, while operating in a regime of high optomechanical-backaction and millikelvin base temperature. This reduced heating translates into a demonstrated phonon-to-photon conversion efficiency of 93.1±0.8%{93.1} \pm {0.8}\% at an added noise of 0.25±0.01{0.25} \pm {0.01} quanta, representing a significant advance toward quantum-limited microwave-optical frequency conversion and optically controlled quantum acoustic memories.


Geometry and operation principle of the graphene-based optical memristor. (a) Schematic concept configuration. Light guided in the waveguide interacts with graphene via its evanescent field. Electrical pulses of the gate voltage ( ${V_G}$ ) are applied to graphene via ferroelectric P(VDF-TrFE) to nonvolatile change graphene’s refractive index ( $n$ ) and absorbance ( $\kappa$ ) to modulate the phase and amplitude of the optical output. (b) Operation mechanism of the graphene-based optical memristor. Nonvolatile variations of graphene’s $n$ and $\kappa$ and waveguide transmission occur in respect to pulsed ${V_G}$ . ${{ P}_{\rm{up}}}$ and ${{P}_{\rm{down}}}$ represent polarization-up and -down states in ferroelectrics after the SET and RESET operations, holding the carrier doping states in graphene for nonvolatile State 1 and State 2. Band structures of graphene with different doping levels are shown on the right side of the inset. (c) Schematic of the looped hysteretic responses of graphene’s $n$ and $k$ and resulting optical transmission states with $V_G$ in the optical memristor.
On-chip graphene amplitude memristor demonstration. (a) Optical microscope image of a fabricated device. The inset shows the full-scale view of the device. (b) Hysteresis of the variations in the measured waveguide transmission and source-drain current ( ${I_{\textit{DS}}}$ ) of the graphene channel for ${V_G}$ cyclically scanning in a range of ${-}{20}$ to 20 V at a specific wavelength of 1550 nm. Black arrow indicates the ${V_G}$ sweep direction. (c) Demonstration of optical memristor operation at sub-millisecond level. The HIGH (LOW) transmission state is converted by ${V_G}$ pulses of ${-}{20}\;{\rm V}$ (20 V) and maintained afterwards. (d) Cyclability of the SET/RESET operation for 400 switching events. The shaded area indicates fluctuations of 0.04 dB for cycling operations. (e) Transmission retention time for two states. Retention time is measured to over 24 h, and the dashed line marks extrapolated transmissions for over 10 years without crossing each other. (f) Broad-spectrum response of graphene-based optical memristors. Twelve switching transitions were performed randomly within 24 h. The shaded area indicates the standard deviation and the solid line indicates the average. (g) Semicontinuous nonvolatile operations of optical memristors. The device is returned to the base state by a RESET pulse of ${V_G} = {20}\;{\rm V}$ . The intermediate state is determined by the amplitude of the SET pulse. The inset zooms into a complete operation of the intermediate state.
On-chip graphene phase memristor demonstration. (a) Top-view optical microscope image of a fabricated graphene-memristor with an MRR. The inset shows an image of the device before the preparation of the gate electrode. (b) Transmission spectra of a graphene-MRR optical memristor in various levels. Three intermediate states are excited by ${-}{14},\;- {13}$ , and ${-}{12}\;{\rm V}$ pulses. Zoomed transmission spectra around the resonant wavelengths of 1531 nm (c) and 1561 nm (d). Hysteresis response of ${Q}$ factors (e) and shift of resonant wavelength (f) with the cyclically scanned ${V_G}$ at three specific resonances. (g), (h) Hysteresis response of transmission as a function of ${V_G}$ at two resonant wavelengths. All curved arrows indicate the direction of the curve’s hysteresis.
Photonic architecture for in-memory matrix-vector multiplication (MVM) based on graphene optical memristor. (a) Optical micrograph of the device systems fabricated to demonstrate computational architectures. Two graphene-MRR memristor cells are evanescently coupled with the same bus-waveguide. By applying input voltage pulses (Input A and Input B) independently to the gates of the two MRRs, the transmittance values of their resonant modes are pre-SET and stored as Ta and Tb. Cooperating with the intensity of the input optical signals (Ia and Ib) at the resonant wavelengths of the two MRRs, individual multiplication is accomplished. A photodetector at the output port of bus-waveguide as an adder of optical signal to obtain the MVM calculation ( ${\rm Ia} \times {\rm Ta} + {\rm Ib} \times {\rm Tb}$ ). (b) Demonstration of a WDM operations of the fabricated device systems. The blue and red resonant dips correspond to the modes in two MRRs. Each cell can be operated individually while another cell is not affected. (c) Demonstration of MVM. The figure is divided into four sections, corresponding to Ta and Tb being nonvolatile reprogrammed to four sets of values. A series of input optical signals is reused in each section to verify the accuracy of the MVM calculations. The measured values are compared with numerically calculated exact values.
An architecture of in-memory logic gate operations based on two cascaded graphene-MRRs. (a) Transmission spectrum near the common resonant wavelength (1586.77 nm) of two graphene-MRRs. Two MRRs at their respective resonant wavelengths (the blue and red shaded areas) can be nonvolatile modulated by their respective VGs. They share a common resonant wavelength (the green shaded area) due to the different free spectral regions (FSR). The cascaded device system exhibits HIGH transmission at the common resonant wavelength only if both MRRs exhibit HIGH transmission (after ${V_{\textit{Ga}}} = {20}\;{\rm V}$ & ${V_{\textit{Gb}}} = {20}\;{\rm V}$ ). (b) After setting the MRRa to two states by applying pulsed voltages of 20 (orange curve) or ${-}{20}\;{\rm V}$ (blue curve) on ${V_{\textit{Ga}}}$ , the hysteresis transmission response of the bus-waveguide with ${V_{\textit{Gb}}}$ at the common resonant wavelength of 1586.77 nm. (c) Optical output transmission states showing AND logic operations, with ( ${V_{\textit{Ga}}},\;{V_{\textit{Gb}}}$ ) as the logic inputs ( $A,\;B$ ), and the bus-waveguide transmission ( $T$ ) as the output.
On-chip optical memristors based on ferroelectric-doped graphene

January 2025

·

37 Reads

Chip-integrated optical memristors, modulating light in a nonvolatile and semicontinuous manner, are attractive to revolutionize on-chip optical signal processing via the constructions of nonvolatile reconfigurable photonic circuits, in-memory computing, brain-inspired architectures, etc. Mechanisms, including phase-change, filamentation, and ferroelectricity, have been attempted to implement on-chip optical memristors, though their intricate tradeoffs between fabrication compatibility, modulation depth, power consumption, retention time, and cyclability make it desired to pursue new architectures. Here, we demonstrate graphene-based on-chip optical amplitude and phase memristors by electrostatically doping the graphene integrated on a silicon nitride waveguide with a ferroelectric film. Benefiting from graphene’s significant dependence of complex refractive index on its carrier density and the ferroelectric remnant doping, semicontinuous nonvolatile modulation with a maximum depth of 32.5  dB{\sim}{32.5}\;{\rm dB} is realized with a low programming energy of {\sim}{1.86}\;{\rm pJ/}\unicode{x00B5}{\rm m}^2 , exhibiting good cyclability (fluctuation ratio <0.9%{\lt}{0.9}\% ) and long retention time (over 10 years). By integrating the graphene-based optical memristor with cascaded microring resonators, in-memory computings with multiple wavelength channels are demonstrated by analogue matrix-vector multiplication and digital logic gate operations. Combining these merits with CMOS-compatible on-chip graphene integration, the demonstrated graphene-based optical memristor has proven to be a competitive candidate for high-bandwidth neuromorphic computing, convolutional processing, and artificial intelligence on photonic integrated circuits.


Accurate vector optically pumped magnetometer with microwave-driven Rabi frequency measurements

January 2025

·

27 Reads

Robust calibration of vector optically pumped magnetometers (OPMs) is a nontrivial task, but increasingly important for applications requiring high-accuracy such as magnetic navigation, geophysics research, and space exploration. Here, we showcase a vector OPM that utilizes Rabi oscillations driven between the hyperfine manifolds of 87Rb^{87}{\rm Rb} to measure the direction of a DC magnetic field against the polarization ellipse structure of a microwave field. By relying solely on atomic measurements—free-induction decay (FID) signals and Rabi measurements across multiple atomic transitions—this sensor can detect drift in the microwave vector reference and compensate for systematic shifts caused by off-resonant driving, nonlinear Zeeman (NLZ) effects, and buffer gas collisions. To facilitate deadzone-free operation, we also introduce a Rabi measurement that utilizes dressed-state resonances that appear during simultaneous Larmor precession and Rabi driving (SPaR). These measurements, performed within a microfabricated vapor cell platform, achieve an average vector accuracy of 0.46 mrad and vector sensitivities down to 11\;{\unicode{x00B5}}{\rm rad}/\sqrt {{\rm Hz}} for geomagnetic field strengths near 50 µT. This performance surpasses the challenging 1-deg (17 mrad) accuracy threshold of several contemporary OPM methods utilizing atomic vapors with an electromagnetic vector reference.


Single-shot spontaneous Raman scattering for time-resolved one-dimensional hypersonic flow diagnostics

January 2025

·

22 Reads

Spontaneous Raman scattering, also known as common old ordinary Raman scattering (COORS), is revisited to evaluate its applicability for hypersonic flow characterization. Due to its very low cross section, Raman scattering is often considered unsuitable for measuring low-pressure gas properties that are found in ground test simulations of high-altitude hypersonic flights. Utilizing a recently developed one-dimensional (1D) light scattering technique with a volume Bragg grating filter and Stokes sideband windowing, we demonstrate 1D rotational Raman measurements of temperature and neutral gas density across a bow shock in front of a blunt wedge model under Mach 6 hypersonic flow. The experiment was conducted in the Hypervelocity Expansion Tunnel at Texas A&M University. The measurements were successfully obtained during a single run of the tunnel operation, capturing the temperature and density distributions with dynamic ranges of 200–2000 K and 5×10234×1024/m35 \times {10^{23}} - 4 \times {10^{24}} /{{\rm m}^3} , respectively, over both the free-stream and post-shock regions, covering approximately 10 mm in length with a spatial resolution of <0.5  mm{\lt}0.5\;{\rm mm} . Time-resolved high-speed measurement capability at 100 kHz was also demonstrated, showcasing the robustness of 1D COORS for gas diagnostics.


Computational optical imaging: on the convergence of physical and digital layers

January 2025

·

34 Reads

Optical imaging has traditionally relied on hardware to fulfill its imaging function, producing output measures that mimic the original objects. Developed separately, digital algorithms enhance or analyze these visual representations, rather than being integral to the imaging process. The emergence of computational optical imaging has blurred the boundary between hardware and algorithm, incorporating computation in silico as an essential step in producing the final image. It provides additional degrees of freedom in system design and enables unconventional capabilities and greater efficiency. This mini-review surveys various perspectives of such interactions between physical and digital layers. It discusses the representative works where dedicated algorithms join the specialized imaging modalities or pipelines to achieve images of unprecedented quality. It also examines the converse scenarios where hardware, such as optical elements and sensors, is engineered to perform image processing, partially or fully replacing computer-based counterparts. Finally, the review highlights the emerging field of end-to-end optimization, where optics and algorithms are co-designed using differentiable models and task-specific loss functions. Together, these advancements provide an overview of the current landscape of computational optical imaging, delineating significant progress while uncovering diverse directions and potential in this rapidly evolving field.


Operation principle of the proposed PCMT device. (a) Structure and polarization modulation principle of the PCMT device. The composition of the liquid–crystal (LC) phase shifter is shown in the dashed gray box. (b) Phase difference between p- and s-reflections ( ${\varphi _{\rm{ps}}}$ ) caused by the mirror–TIR structure as a function of prism–mirror distance $h$ at 1, 2, and 3 THz, respectively. (c) In-plane birefringence $\Delta n$ of the LC as a function of applied voltage $V$ . (d) ${\varphi _{\rm{ps}}}$ by the mirror–TIR structure, LC phase shifter, and the combined PCMT device at four combinations of $h$ and $\Delta n$ . Arrows and circles indicate the approximated polarization states by the combined PCMT device in 1.5–3.5, assuming a 45° linearly polarized incidence.
Phase control by the mirror–TIR component. (a) Phase difference ( ${\varphi _{\rm{ps}}}$ ) and (b) amplitude ratio ( $| {{r_p}/{r_s}} |$ ) between p- and s-reflections at 11 prism–mirror distance $h$ . Circles are the experimental results. Solid curves in (a) are the model fits and dashed gray line in (b) gives the theoretical unity ratio. (c) Evolution of the time-domain waveforms of ${E_p}$ and ${E_s}$ with increasing $h$ (offset for clarity). The same color scheme applies to (a–c), indicated by the arrows in (a) and (c). (d) Comparison of the set $h$ and the best fit for the 11 measurements.
Characterization and active control of the liquid-crystal (LC) phase shifter. (a) Refractive indices $n$ , birefringence $\Delta n$ , and (b) absorption coefficients $\alpha$ of the used LC NJU-LDn-4. (c) Time response of $\Delta n$ to the ON and OFF bias switch. The inset zooms in the rapid drop of $\Delta n$ after bias ON. (d) Hysteresis response of $\Delta n$ to bias voltage modulated by square waves with different duty cycles ( $D$ ). Forward and backward correspond to increasing and decreasing $D$ , respectively. The inset shows the full 0–100% response.
Multifunctional polarization control of the integrated device. (a) Phase differences ( ${\varphi _{\rm{ps}}}$ ) and (b) amplitude ratios ( $| {{r_p}/{r_s}} |$ ) between p- and s-reflections at the optimized settings of the four polarization states. (c) Time evolution of the electric field corresponding to the four polarization states in 1.6–3.4 THz, stepped by 0.2 THz and represented by different colors indicated in the color bar. The arrows point to the rotational directions. DoLP and DoCP refer to the degree of linear/circular polarization, respectively.
Conversions of arbitrary polarization states centered at customizable frequencies. (a) Phase differences ( ${\varphi _{\rm{ps}}}$ ) between p- and s-reflections for the three targeted polarization states, including six alternative settings. Symbols are the experimental results and solid curves are the theoretical predictions. Gray areas represent the range with $\pm {10}^\circ$ to the target phase. (b–g) Time-evolution of the electric field for the six groups of settings. Colors represent different frequencies in units of THz. Dotted black curves indicate the targeted output. Arrows point to the rotational directions. The rectangle in (b) illustrates the definition of the $(\theta ,\;\varepsilon)$ coordinates.
Achromatic arbitrary polarization control in the terahertz band by tunable phase compensation

January 2025

·

53 Reads

Polarization is a key parameter in light–matter interactions and is consequently closely linked to light manipulation, detection, and analysis. Terahertz (THz) waves, characterized by their broad bandwidth and long wavelength, pose significant challenges to efficient polarization control with existing technologies. Here, we leverage the mesoscale wavelength characteristics of THz waves and employ a mirror-coupled total internal reflection structure to mechanically modulate the phase difference between p- and s-waves by up to 289°. By incorporating a liquid crystal phase shifter to provide adaptive phase compensation, dispersion is eliminated across a broad bandwidth. We demonstrate active switching of orthogonal linear polarizations and handedness-selective quarter-wave conversions in the 1.6–3.4 THz range, achieving an average degree of linear/circular polarization exceeding 0.996. Furthermore, arbitrary polarization at any center frequency is achieved with a fractional bandwidth exceeding 90%. This customizable-bandwidth and multifunctional device offers an accurate and universal polarization control solution for various THz systems, paving the way for numerous polarization-sensitive applications.


Illustration of a Mach–Zehnder interferometer that includes 50:50 nonpolarizing beam splitters (BS), an optical element described by a unitary matrix ${{\textbf M}}$ , and two mirrors. A random light beam ${{\textbf E}_0}$ is divided into two arms, and in arm 2 the element ${\textbf M}$ introduces a cyclic polarization-state change. Interference of the fields ${{\textbf E}_1}$ and ${{\textbf E}_2}$ from the arms is considered at the output plane $z = 0$ some distance away from the last beam splitter. In the example of Section 3.C and in the experiments of Section 4 ${\textbf M}$ is composed of five waveplates.
Behavior of the Poincaré vectors ${\hat{\textbf s}}$ , ${{\hat{\textbf s}}_ +}$ , and ${{\hat{\textbf s}}_ -}$ in a cyclic, unitary polarization-state evolution within the unit Poincaré sphere in the Stokes parameter ( ${s_1},{s_2}$ , ${s_3}$ ) space. The tip of the vector ${\hat{\textbf s}}$ is on the inner ball whose radius is the degree of polarization $P$ , whereas ${{\hat{\textbf s}}_ \pm}$ correspond to orthogonal fully polarized states on the surface of the unit-radius sphere. The solid angles of the paths traversed by the three vectors are the same and denoted by $\Omega$ .
Illustration of the closed paths on the Poincaré sphere related to the evolution of (a) $y$ -polarized ( ${{\hat{\textbf u}}_ +}$ ) and (b) $x$ -polarized ( ${{\hat{\textbf u}}_ -}$ ) components of a partially polarized beam passing through the sequence of waveplates described in the main text. The north and south poles correspond to right-hand circular and left-hand circular polarization states, respectively, while the points on the equator refer to the linearly polarized states at an angle of $2\theta$ to the $x$ axis. In the experiments of Section 4, the blue-highlighted points on the equator are moved.
Schematic illustration of the setup for the measurement of phase-difference variations of partially polarized beams. The preparation stage produces a beam composed of uncorrelated $x$ and $y$ components with the latter having a larger intensity. The Mach–Zehnder interferometer corresponds to that in Fig. 1 and the five waveplates ${{\textbf M}_1} \ldots {{\textbf M}_5}$ are described in Appendix C. C: collimator and aperture; DL: delay line; PBS: polarizing beam splitter; ND: neutral density filter; BS: 50:50 beam splitter; QWP: quarter-wave plate; HWP: half-wave plate.
Behavior of the measured (blue circles) and theoretical (solid line) phase changes $\Delta {\varphi _{\text{P}}}$ as a function of the angle $2\theta$ for the degree of polarization values of (a) $P \approx 1$ (full polarization), (b) $P \approx 0.75$ , (c) $P \approx 0.50$ , (d) $P \approx 0.25$ . The quantity $2\theta$ is the angle of linear polarization related to the point highlighted in Fig. 3(a) in blue and the bars show the errors based on standard deviations.
Pancharatnam’s connection and geometric phase of partially polarized light

January 2025

·

64 Reads

We extend the concepts of Pancharatnam’s connection and phase difference of optical beams to cover random, stationary, partially polarized quasimonochromatic light. The geometric and dynamic structures of the phase are assessed within the framework of classical statistical polarization theory when the beam undergoes a cyclic, unitary polarization-state evolution. The existence of the two phase types is confirmed experimentally by conducting an interferometric measurement where the phase changes are found from the variations in the position of an intensity fringe pattern. The formalism also leads to a definition of the geometric phase for partially polarized beams, which is consistent with the results on mixed (qubit) quantum states.


Fiber design and optical property simulations. Cross sections of (a) the 5-DNANF, (b) the 4T-DNANF, and (c) the 4T-DNANF with larger gap, with their structural parameters listed aside. (d), (e), (f) FEM-simulated CLs of the ${{\rm LP}_{01}}$ (black), ${{\rm LP}_{11}}$ (red) core modes, and HOMER (blue) as a function of ${Z_1}/{R_{{\rm core}}}$ and ${Z_2}/{R_{{\rm core}}}$ for (a), (b), and (c), respectively, where ${Z_1}$ and ${Z_2}$ are the thicknesses of the first and second air layer in the cladding and ${R_{{\rm core}}}$ is the core radius. (g) Minimum achievable CL for the structures in (a) and (b) as a function of gap size (black curve, with core radius and glass wall thickness fixed) and core radius (blue curve, with glass wall thickness fixed and gap size linearly scaled), respectively. (h) Effective indices of the core ${{\rm LP}_{01}}/{{\rm LP}_{11}}$ modes (black/red), the first cladding air crescent ${{\rm LP}_{01}}$ mode (blue), and the second cladding air crescent ${{\rm LP}_{01}}$ mode (green), along with their corresponding mode-intensity profiles ( $|E|$ ) in linear scale. The plots inside the green and blue dashed squares show the intensity ( $|E|$ ) of the core and cladding modes on a logarithmic scale, respectively. (i) Simulated CL and HOMER as a function of wavelength at two ${Z_1}/{R_{{\rm core}}}$ values of 0.65 and 0.88, using the structural parameters of (c).
AR-HCF fabrications and loss measurements. (a)–(e) SEM images of the five fabricated 4T-DNANFs with their structural parameters listed in (l). Scale bar, 50 µm. (f) Measured transmission (blue and red) and loss (black) spectra of Fiber #1. The measurement uncertainty is indicated by the grey shaded area. (g) Loss spectrum of Fiber #1 measured with a wavelength resolution of 0.05 nm, revealing the absorption lines from ${{\rm CO}_2}$ gas. (h) one-way OTDR traces of a 4T-DNANF (Fiber #1), a G.654E fiber, and a G.652D fiber in the left axis, and accumulated losses measured by two-way OTDR in the right axis. (i) Same as (f) for Fiber #5. (j) ${{\rm S}^2}$ -imaging measurement of Fiber #5 with varying fiber lengths. The insets show the reconstructed mode-intensity profiles at two different differential group delays. (k) Fitted loss curve of the ${{\rm LP}_{11}}$ mode derived from (j) for Fiber #5. (l) Summaries of structural parameters, modal losses, and HOMER for the five AR-HCFs. The variations in structural parameters are obtained by measuring the cross-section at both ends of the several-km-long fiber. (m) Measured average losses of the ${{\rm LP}_{01}}$ and ${{\rm LP}_{11}}$ modes in the 1540–1560 nm range, along with the corresponding HOMER for Fibers #1 to #5.
Fourfold truncated double-nested antiresonant hollow-core fiber with ultralow loss and ultrahigh mode purity

January 2025

·

48 Reads

Hollow-core fibers (HCFs) are inherently multimode, making it crucial to filter out higher-order modes (HOMs) within the shortest possible fiber length for applications such as high-speed coherent communications and fiber-optic gyroscopes. However, current HCF designs face the challenges of simultaneously achieving ultralow fundamental mode (FM) loss and ultrahigh HOM suppression. In this study, we present a fourfold truncated double-nested antiresonant nodeless hollow-core fiber (4T-DNANF) structure that addresses this challenge. Our 4T-DNANF enables greater control over phase matching between core modes and air modes in the cladding, allowing for minimized FM loss and substantially increased HOM loss. Experimentally, we fabricated several HCFs: one with an FM loss of 0.1 dB/km and an HOM loss of 430 dB/km, and another achieving an FM loss of 0.13 dB/km with a HOM loss of 6500 dB/km, yielding a higher-order mode extinction ratio of 5×104{5} \times {{10}^4} —the highest reported to date.






Overview of multifunctional photonic memory concept. (a) Illustration and working mechanism of multifunctional photonic memory leveraging a PN junction microheater and modulator. Coarse nonvolatile tuning of the PCM is achieved by applying a forward-biased pulse to locally heat the waveguide while reverse biasing the PN junction modulates carriers in the depletion region for volatile fine-tuning. (b) Measured current density as a function of voltage applied to the waveguide-integrated PN junction. Biasing the PN junction below ${\sim}{1.5}\;{\rm V}$ enables volatile tuning without significant heating of the PCM. (c) Modeling comparison of heat generated in PIN, PN, and ${ n}$ -doped waveguide-integrated microheaters. All simulations use a fixed ${+}{5}\;{\rm V}$ forward bias and 1 µm spacing between ${n+}\!+ /{p+}\!+$ regions. (d) Optical microscope image of a fabricated microring array (bottom, 25 µm scale bars) and magnified image of a single microring with integrated multifunctional memory cell (top left, 10 µm scale bar). Cross-sectional view of a PN microheater with PCM deposited in the oxide window (top right).
Effects of annealing step on deposited PCM ( ${{\rm Ge}_2}{{\rm Sb}_2}{{\rm Se}_1}{{\rm Te}_4}$ ). (a) Observed initial spectral blue shift prior to reaching the fully crystallized state due to reflow of the as-deposited PCM above the glass transition temperature. (b) After reflow, reversible switching is achieved, demonstrating a topographical change in the PCM rather than a structural or chemical change. (c) Observed spectral red shift of both resonances after annealing the entire chip on a hot plate. The large red shift is expected and indicates minimal reflow of the PCM after annealing. (d) Spectrum of the microring with reamorphized PCM matches well with the initial as-deposited state, demonstrating minimal topographical changes after hot plate annealing.
Nonvolatile tuning of photonic memory. (a),(b) Nonvolatile programming of arbitrary optical weights using memory cell arrays with four MRRs containing (a) GSST and (b) ${\rm Sb}_2{\rm Se}_3$ . For these experiments, Ring 1 was fully amorphized, Rings 2 and 3 were partially amorphized, and Ring 4 was left in the crystalline state. (c) Incremental amorphization of GSST using a PN microheater (probe wavelength indicated by the dashed line). (d) Measured blue shift and (e) transmission state of Ring 1 from (c) as a function of amorphization pulse number. Amorphization pulses were 100 µs in width with an increasing voltage ranging from 5.6 V to 6.4 V.
Volatile tuning of photonic weight. (a) Measured resonance shift as a function of reverse bias voltage. The red line shows the expected resonance shift based on the analytical model of carrier modulation in the PN depletion region. (b) Shift in MRR resonance as a function of forward bias voltage. After ${\sim}{1.5}\;{\rm V}$ , Joule heating begins to dominate and shifts the resonance back to longer wavelengths. (c) Static power consumption versus resonance shift for the PN junction in both forward and reverse bias configurations. Reverse biasing the device provides ${\gt}{10}{,}{000} \times$ better power efficiency. (d) Normalized frequency-dependent electro-optic response of the PN junction in both forward and reverse bias conditions. Carrier recombination limits the frequency response to ${\sim}{200}\;{\rm MHz}$ under forward bias, while the RF bandwidth of the measurement setup limits the response in reverse bias.
Multifunctional operation. (a) Demonstration of both nonvolatile and volatile tunability in our multifunctional memory cell. The PCM is cycled between partially amorphous and fully crystalline states with increasing amorphization pulse amplitude. Blue regions show volatile tuning of the PN junction before returning the PCM to the crystalline state. (b) Zoomed-in region from (a) showing both volatile and nonvolatile tunability. Small discontinuities indicated by vertical arrows are due to noise from electrical relays used to switch between instruments. (c) Nonvolatile cyclability of our memory cell between the fully amorphous and fully crystalline states. The resonance of the ring in the crystalline state blue shifts during cycling (see inset), indicating the endurance is limited by a degradation of the crystalline state.
High-speed multifunctional photonic memory on a foundry-processed photonic platform

January 2025

·

33 Reads

The integration of computing with memory is essential for distributed, massively parallel, and adaptive architectures such as neural networks in artificial intelligence (AI). Accelerating AI can be achieved through photonic computing, but it requires nonvolatile photonic memory capable of rapid updates during on-chip training sessions or when new information becomes available during deployment. Phase-change materials (PCMs) are promising for providing compact, nonvolatile optical weighting; however, they face limitations in terms of bit precision, programming speed, and cycling endurance. Here, we propose a novel photonic memory cell that merges nonvolatile photonic weighting using PCMs with high-speed, volatile tuning enabled by an integrated PN junction. Our experiments demonstrate that the same PN modulator, fabricated via a foundry-compatible process, can achieve dual functionality. It supports coarse programmability for setting initial optical weights and facilitates high-speed fine-tuning to adjust these weights dynamically. The result shows a 400-fold increase in volatile tuning speed and a 10,000-fold enhancement in efficiency. This multifunctional photonic memory with volatile and nonvolatile capabilities could significantly advance the performance and versatility of photonic memory cells, providing robust solutions for dynamic computing environments.


Noise2Image: noise-enabled static scene recovery for event cameras

January 2025

·

54 Reads

·

1 Citation

Event cameras, also known as dynamic vision sensors, are an emerging modality for measuring fast dynamics asynchronously. Event cameras capture changes of log-intensity over time as a stream of “events” and generally cannot measure intensity itself; hence, they are only used for imaging dynamic scenes. However, fluctuations due to random photon arrival inevitably trigger noise events, even for static scenes. While previous efforts have been focused on filtering out these undesirable noise events to improve signal quality, we find that, in the photon-noise regime, these noise events are correlated with the static scene intensity. We analyze the noise event generation and model its relationship to illuminance. Based on this understanding, we propose a method, called Noise2Image, to leverage the illuminance-dependent noise characteristics to recover the static parts of a scene, which are otherwise invisible to event cameras. We experimentally collect a dataset of noise events on static scenes to train and validate Noise2Image. Our results show that Noise2Image can robustly recover intensity images solely from noise events, providing an approach for capturing static scenes in event cameras, without additional hardware.


Wave propagation between two non-parallel planes. (a) Coordinate transformation, where the $uv$ plane can be viewed as the $xy$ plane that undergoes two successive rotations around the $z$ -axis by an angle of $\phi$ , and then $v$ -axis by an angle of $\theta$ . (b) Principles of our method.
Simulation results of scalar diffraction. (a) Amplitude and phase distribution on the source plane. (b) Calculated intensity and phase of the diffractive fields of GT (left column), our method (middle column), and the Ctrl method (right column). Note that a common linear phase term, induced by the intersection angle, is removed for visualization. (c) Field deviations, computed by the absolute value of the direct subtraction, between the two methods and GT. (d) Mean error under different incident fields of the two methods as a function of intersection angles. The error function is defined by Supplement 1, Eq. (S17). (e) Volumetric intensity distribution of the diffractive field, where the observation plane is colored in blue. (f) Time consumption as a function of errors. Note that the errors in (d) and (f) are shown in logarithmic values, since our method is far more accurate than the Ctrl one.
Simulation results of vectorial diffraction. (a) Polarization, amplitude, and phase distribution on the pupil plane. The pseudo color for polarization visualization denotes ellipticity. (b) Calculated intensity and phase of the $x$ -, $y$ -, $z$ -polarized component, and the total intensity, of the diffractive fields of GT (top row), our method (middle row), and the Ctrl method (bottom row). (c) Volumetric intensity distribution of the diffractive field, where the observation plane is colored in blue. (d) Field deviations between the two methods and GT of the $x$ -, $y$ -, $z$ -polarized components and the total intensity. (e) Mean error under different pupils of the two methods as a function of intersection angles. (f) Time consumption as a function of errors.
CGH on a tilted plane. (a) Simplified schematic of the optical setup. The black arrow represents the incident direction of the laser. The bottom right patterns are the holograms calculated from our method and the Ctrl one, respectively; L1–L2, lenses. (b) Generated patterns utilizing our method (left column) and the Ctrl one (right column) in the simulation. (c) Experimental results. (d) Volumetric maximum intensity projection, revealing the highest intensity structure within a volume, of diffractive field calculated from our method (left column) and the Ctrl one (right column) in the experiment.
Diffraction modeling between arbitrary non-parallel planes using angular spectrum rearrangement

January 2025

·

127 Reads

Numerical modeling of diffraction between tilted planes provides remarkable flexibility in computational optics, enabling convenient prediction and manipulation of light on complicated geometries. Specifically it enables, for example, efficient simulation of wave propagation through lenses, fast calculation of holograms for meshed three-dimensional objects, and trapping particles in complicated shapes. However, computational accuracy and efficiency of existing methods are often at odds with each other. Here, we present an approach that accurately and efficiently models wave propagation between two arbitrary non-parallel planes, which is realized by rearranging the angular spectrum of the source field, coupled with a Fourier transform algorithm that does not require zero-padding and uniform sampling. It applies to both scalar and vectorial diffraction modeling, achieving a 1010910 - {10^9} times accuracy improvement, depending on different intersection angles. Notably, our method can cope well with orthogonal-plane diffraction, which is inaccessible in previous methods. Moreover, it enables a flexible balance between accuracy and efficiency, providing potential for further acceleration and accuracy enhancement. After theoretical verification, we provide experimental demonstration in computer-generated holography.



Measurement principle of coherently controlled quartz-enhanced photoacoustic spectroscopy (COCO-QEPAS). (a) Setup scheme for COCO-QEPAS. The tunable laser light of a fiber-feedback optical parametric oscillator (FFOPO) is modulated by means of an acousto-optic modulator (AOM). At absorptive wavenumbers ${\nu _{{\rm vib,rot}}}$ , this excites the fork at its modulation frequency. A coherent control pulse sequence is phase-shifted by $\Delta \phi$ . (b) Residual oscillation of the fork due to its high Q-factor after excitation. Coherent control allows deceleration of the fork. First, the fork is accelerated (Excitation). Then, the oscillation amplitude reveals spectral information (Measurement). Eventually, the fork is actively damped (Coherent control), enabling a fast next measurement.
Residual signal dependent on the phase shift between excitation and damping sequence. (a) Modulated laser power. The phase shift between excitation and damping cycles is color-coded in blue (0) to red ( $2\pi$ ). (b) Time trace of the demodulated cell signal for exciting with a sequence of 50 cycles and phase shifts of 50 damping cycles from 0 to $\pi$ . The remaining residual cell signal at 10 ms is minimal for a phase shift of $\pi$ . (c) Demodulated cell signal for further increasing the phase shift to $2\pi$ . In (d) the dependence of the residual signal on the phase shift is plotted. No residual signal is left at $\pi$ .
Comparison of COCO-QEPAS with conventional QEPAS. (a) Acquisition of ${\rm CH}_{4}$ spectra at increasing sweeping speeds from 5 to 125 nm/s. The fork relaxation time reflects in increasingly smeared out spectral features at higher sweeping speeds. COCO-QEPAS overcomes this detrimental effect. (b) Quantification of improved spectral quality in coherent control by the quadratic deviation from the convolved HITRAN data. (c) Fastest COCO-QEPAS sweep in three seconds (orange) overlapped with HITRAN data. The spectral features are still visible for all branches.
Coherent control in quartz-enhanced photoacoustics: fingerprinting a trace gas at ppm-level within seconds

January 2025

·

23 Reads

Quartz-enhanced photoacoustic spectroscopy (QEPAS) has become a versatile tool for detection of trace gases at extremely low concentrations, leveraging the high quality (Q)-factor of quartz tuning forks. However, this high Q-factor imposes an intrinsic spectral resolution limit for fast wavelength sweeping with tunable laser sources due to the long ringing time of the tuning fork. Here, we introduce a technique to coherently control and damp the tuning fork by phase-shifting the modulation sequences of the driving laser. Particularly, we send additional laser pulses into the photoacoustic cell with a timing that corresponds to a π\pi phase shift with respect to the tuning fork oscillation, effectively stopping its oscillatory motion. This enables acquisition of a complete methane spectrum spanning 3050–3450 nm in just three seconds, preserving the spectral shape. Our measured data is in good agreement with the theoretically expected spectra from the HITRAN database when convolved with the laser linewidth of <2  cm1{\lt}2\;{{\rm cm}^{- 1}} . This will leverage the use of QEPAS with fast-sweeping OPOs in real-world gas sensing applications beyond laboratory environments with extremely fast acquisition speed enabled by our coherent control scheme.


Mouse abdomen before (a) and after applying 4-aminoantipyrine (b) and tartrazine (c). Dashed green box in (a) indicates region in which OCT and PAM imaging were performed. Scale bar: 1 cm.
Comparison of OCT B-scan images in pigmented vs. nonpigmented mouse using 4-aminoantipyrine and tartrazine. (a)–(c) OCT B-scan image of pigmented mouse abdomen before (a) and after treatment with 4-aminoantipyrine (b) and tartrazine (c). (d)–(f) OCT B-scan image of nonpigmented mouse abdomen before (d) and after treatment with 4-aminoantipyrine (e) and tartrazine (f). Scale bars: 500 µm. Yellow arrows indicate the surface of the topical tissue clearing solution.
Histological comparison of OCT images. (a) OCT B-scan image of nonpigmented mouse skin after applying tartrazine. (b) Enlarged view of region enclosed by dashed green box in (a). (c) H&E-stained paraffin-embedded abdominal skin sample from wild type BALB/c mouse [18]. (d) OCT B-scan image of pigmented mouse treated with 4-aminoantipyrine. (e) Enlarged view of region enclosed by dashed orange box in (d). Yellow arrows denote hair follicles beneath the skin surface. (f) Enlarged view of region enclosed by dashed blue box in (d). Yellow arrows denote possible adipocyte nuclei located in the intradermal fat layer. Scale bars: 500 µm (yellow), 100 µm (magenta and black).
OCT time course imaging of pigmented mouse abdominal skin before and after topical application of 4-aminoantipyrine. (a)–(g) OCT B-scan images of mouse abdominal skin before (a) and 1 min (b), 5 min (c), 10 min (d), 15 min (e), 20 min (f), and 30 min (g) following dye application. (h) Plot of intensity line profile extracted from the dashed red line in (a) and dashed green line in (f). Scale bars: 500 µm. Yellow arrows indicate residual grains of pumice sand.
PA imaging results. (a, b) Maximum amplitude projection (MAP) of PA images of mouse abdomen before (a) and after (b) applying 4-aminoantipyrine solution. ( ${{a}_1}$ ) and ( ${{ b}_1}$ ) are photographs of the scanning region for (b) and (c), correspondingly. The white arrow indicates a matching blood vessel in (a) and (b). (c, d) Cross-sectional PA images before (c) and after (d) application of dye solution corresponding to the dashed lines in (a) and (b), respectively. (e) Line profile extracted from the dashed lines in (a) and (b), L1 and L2, respectively. (f, g) MAP of PA images of mouse abdomen before (f) and after (g) applying tartrazine solution. ( ${{ f}_1}$ ) and ( ${{g}_1}$ ) are photographs of the scanning region for (f) and (g), correspondingly. (h, i) Cross-sectional PA images before (h) and after (i) application of dye solution corresponding to the dash lines in (f) and (g), respectively. (j) Line profile extracted from the dashed lines in (f) and (g), L3 and L4, respectively. Scale bars: 1 mm (white), 0.5 mm (yellow).
Enhanced penetration depth in optical coherence tomography and photoacoustic microscopy in vivo enabled by absorbing dye molecules

January 2025

·

84 Reads

The scattering and absorption of light within biological tissue severely limits the penetration depth of optical imaging techniques. Recently, it has been found that water-soluble, strongly absorbing dye molecules, such as tartrazine, can achieve in vivo tissue transparency by increasing the refractive index of aqueous components in tissue, as predicted by the Lorentz oscillator model and Kramers–Kronig relations. In this study, we topically applied absorbing dye molecules to the abdominal skin of pigmented and nonpigmented mice to enhance the penetration depth of optical coherence tomography (OCT) and photoacoustic microscopy (PAM). In both types of mice, the penetration depth of OCT was significantly improved using tartrazine and 4-aminoantipyrine. As predicted by the Kramers–Kronig relations and absorption spectra of the dyes, mice treated with 4-aminoantipyrine showed significantly improved penetration depth compared to mice treated with tartrazine for the PAM system with 532 nm excitation. These findings further demonstrate the use of absorbing dye molecules for achieving tissue transparency to enhance the penetration depth of depth-resolved optical imaging modalities in skin, thus accelerating the translation of these technologies in clinical areas, such as dermatology.


Optimization of ${{\rm N}_2}$ HCF pressure allows broadening and compression of ps pulses to $\approx 120\;{\rm fs} $ across a wide range of high pulse energies at high average power. ${{\rm N}_2}$ pressure is optimized for maximum broadening without transmission loss. Dispersion was optimized for pulse compression at each repetition rate.
(a) High average and peak power broadening in ${{\rm N}_2}$ for different repetition rates and pulse energies demonstrating $\approx 120\;{\rm fs} $ pulses at 218 W of input power with ${\gt}{70}\%$ transmission efficiency and a constant pressure of 1.5 bar. At repetition rates between 25 and 75 kHz temporal breakdown limits the maximum average power and occurs at 3 mJ of energy per pulse (red dotted line). Inset: SHG-FROG trace for 218 W, 200 kHz at 2.5 bar ${{\rm N}_2}$ starting from 1.06 ps ( ${\tau _{\rm{In}}}$ ) to 126 fs ( ${\tau _{\rm{Ret}}}$ ) FWHM of the intensity. (b) HCF transmission at 25, 50, and 100 kHz repetition rates as a function of input power limited by temporal breakdown or maximum energy reached. (c) Spectral broadening comparison of 218 W input power at 100 kHz in ${{\rm N}_2}$ , Argon (Ar) with 1.5 bar, and Krypton (Kr) at 1.7 bar.
Comparison transmission efficiency and pulse broadening in gasses ${\rm N_2}$ , ${\rm O_2}$ , and ${\rm N_2}{\rm O}$ for 100 kHz, 200 kHz, and 1 MHz repetition rates. (a) Transmission as a function of input power. (b) Transmission vs log of pulse energy. Legend is on the top right. (c) Symmetric spectra from ${\rm N_2}$ at 136 W and 245 W. Largest symmetric and asymmetric spectra generated in ${\rm O_2}$ at 81 W, 136 W and in ${\rm N_2}{\rm O}$ at 27 W, 54 W, respectively, with light/dark spectra for high/low power. Black spectra are that of the input pulse.
(a) Rotational-vibrational energy levels for molecular gasses ${\rm N_2}{\rm O}$ , ${\rm O_2}$ , and ${\rm N_2}$ . First 20 odd number molecular rotational states (colored horizontal lines) superimposed on vibrational states (wider black lines). Only odd rotational states are plotted as Raman selection rules for linear molecules require $\Delta J = \pm 2$ . The population distribution of each gas at room temperature (300 K) is displayed vertically on the left of every gas’s density of states. Energy from 1/2 peak population at room temperature to first vibrational state is given. (b) Bandwidth edges at 1/10 max intensity of pulses broadened HCF with 1.5 bar of ${\rm N_2}$ (red), ${\rm O_2}$ (yellow), and ${\rm N_2}{\rm O}$ (blue) are shown as a function of power. Spectra corresponding to the largest ro-vibrational energy transition gap $\Delta \epsilon$ in each molecule are displayed as the power at $\Delta \epsilon$ bandwidth is reached.
High-harmonic-generation spectra from 200 kHz pulses shown in the FROG trace of Fig. 2(a).
High-power femtosecond molecular broadening and the effects of ro-vibrational coupling

January 2025

·

29 Reads

Scaling spectral broadening to higher pulse energies and average powers is a critical step in ultrafast science, especially for narrowband Yb-based solid state lasers, which have become the new state-of-the-art. Despite their high nonlinearity, molecular gases as a broadening medium inside hollow-core fibers have been limited to 25 W, at best. We demonstrate spectral broadening in nitrogen at ten-fold average powers up to 250 W with repetition rates from 25 to 200 kHz. The observed ten-fold spectral broadening is stronger compared to the more expensive krypton gas and enables pulse compression from 1.3 ps to 120 fs. We identified an intuitive explanation for the observed average power scaling based on the density of molecular ro-vibrational states of Raman active molecules. To verify this ansatz, the spectral broadening limitations in O2{{\rm O}_2} and N2O{{\rm N}_2}{\rm O} are experimentally measured and agree well. On these grounds, we propose a perspective on the role, suitability, and limits of stimulated Raman scattering at high average and peak powers. Finally, high-harmonic generation is demonstrated at 200 kHz. These findings can have strong implications for intense, high-repetition-rate, pulsed ps laser propagation in the atmosphere where the dominant species are N2{{\rm N}_2} and O2{{\rm O}_2} .



Tandem Raman microscopy: integration of spontaneous and coherent Raman scattering offering data fusion analysis to improve optical biosensing

January 2025

·

44 Reads

We provide tandem Raman microscopy (TRAM), a cutting-edge multimodal microscope that integrates the methods of stimulated Raman scattering (SRS), coherent anti-Stokes Raman scattering (CARS), and spontaneous (resonance) Raman scattering [(R)RS]. The device facilitates sequential continuous wave (CW)-driven RS imaging to collect full spectra from every sample location and rapid pulsed-wave-driven SRS-CARS scanning at specific wavenumbers, offering a reliable and efficient analytical tool. The fingerprint spectral region can be included in the spectral imaging capabilities of CARS and SRS. Data collected from a sample area using several techniques can be integrated and analyzed, significantly increasing reliability and predictions. We analyzed the in vitro model of nonadherent leukocytes (LCs) to illustrate the capabilities of this unique system, emphasizing the benefits of measuring the same sample with three different Raman techniques without having to transfer it between microscopes. Data fusion allowed for the correct classification of two subtypes of LCs based on the partial least squares (PLS) discrimination, increasing the prediction accuracy from approximately 83% in the case of textural and morphological data (SRS) to 100% when combined with spectral data (SRS and RS). We also present RRS images of LC labeled with astaxanthin, and reference data from SRS and CARS microscopy. Additionally, polystyrene beads were investigated as a non-biological material. The advantages of each Raman technique are utilized when (R)RS, SRS, and CARS are combined into a single device. This paves the way for dependable chemical characterization in a wide range of scientific and industrial fields.



Journal metrics


10.4 (2023)

Journal Impact Factor™


10 days

Submission to first decision


129 days

Submission to publication


3.669 (2023)

Article Influence Score


Letters & Memoranda USD$2690; Research articles US$3200; $145/page over 8; CC-BY add-on USD$250

Article processing charge

Editors