[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (176)

Search Parameters:
Keywords = whispering gallery mode

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
52 pages, 9743 KiB  
Review
Principles and Applications of ZnO Nanomaterials in Optical Biosensors and ZnO Nanomaterial-Enhanced Biodetection
by Marion Ryan C. Sytu and Jong-In Hahm
Biosensors 2024, 14(10), 480; https://doi.org/10.3390/bios14100480 - 6 Oct 2024
Viewed by 620
Abstract
Significant research accomplishments have been made so far for the development and application of ZnO nanomaterials in enhanced optical biodetection. The unparalleled optical properties of ZnO nanomaterials and their reduced dimensionality have been successfully exploited to push the limits of conventional optical biosensors [...] Read more.
Significant research accomplishments have been made so far for the development and application of ZnO nanomaterials in enhanced optical biodetection. The unparalleled optical properties of ZnO nanomaterials and their reduced dimensionality have been successfully exploited to push the limits of conventional optical biosensors and optical biodetection platforms for a wide range of bioanalytes. ZnO nanomaterial-enabled advancements in optical biosensors have been demonstrated to improve key sensor performance characteristics such as the limit of detection and dynamic range. In addition, all nanomaterial forms of ZnO, ranging from 0-dimensional (0D) and 1D to 2D nanostructures, have been proven to be useful, ensuring their versatile fabrication into functional biosensors. The employment of ZnO as an essential biosensing element has been assessed not only for ensembles but also for individual nanomaterials, which is advantageous for the realization of high miniaturization and minimal invasiveness in biosensors and biodevices. Moreover, the nanomaterials’ incorporations into biosensors have been shown to be useful and functional for a variety of optical detection modes, such as absorption, colorimetry, fluorescence, near-band-edge emission, deep-level emission, chemiluminescence, surface evanescent wave, whispering gallery mode, lossy-mode resonance, surface plasmon resonance, and surface-enhanced Raman scattering. The detection capabilities of these ZnO nanomaterial-based optical biosensors demonstrated so far are highly encouraging and, in some cases, permit quantitative analyses of ultra-trace level bioanalytes that cannot be measured by other means. Hence, steady research endeavors are expected in this burgeoning field, whose scientific and technological impacts will grow immensely in the future. This review provides a timely and much needed review of the research efforts made in the field of ZnO nanomaterial-based optical biosensors in a comprehensive and systematic manner. The topical discussions in this review are organized by the different modes of optical detection listed above and further grouped by the dimensionality of the ZnO nanostructures used in biosensors. Following an overview of a given optical detection mode, the unique properties of ZnO nanomaterials critical to enhanced biodetection are presented in detail. Subsequently, specific biosensing applications of ZnO nanomaterials are discussed for ~40 different bioanalytes, and the important roles that the ZnO nanomaterials play in bioanalyte detection are also identified. Full article
(This article belongs to the Special Issue Low-Dimensional Materials (LDMs) for Biosensing Applications)
Show Figures

Figure 1

Figure 1
<p>An example of a ZnO nanomaterial-based colorimetric biosensor is shown. ZnO NR-modified μPADs were developed as colorimetric biosensors for the detection of glucose and uric acid while employing the chromogenic reagents of AB and AT. (<b>A</b>) The scanning electron microscopy (SEM) panels display the ZnO NRs grown to ~5 μm in length on a Whatman filter paper (WFP). The diameters of the NRs on the μPADs are 500–700 nm for the ZNR-1/WFP, ~200 nm for ZNR-2/WFP, and ~20 nm for ZNR-3/WFP samples. (<b>B</b>) The plot displays the color differences (ΔE) observed from an unmodified WFP, as well as the three ZNR/WFP samples, when 1 mmol/L of uric acid was added to the μPAD sensor containing AB. (<b>C</b>) The colors developed on the ZnO NR-modified μPAD sensors at different incubation times are shown for the detection of both glucose and uric acid. For both analytes, the color images correspond to the case of 1 mmol/L of analyte with AB or AT reacted up to 7 min, as specified for each image. (<b>D</b>) The marked images show the detection layer of a ZNR-3/WFP device when the different concentrations of uric acid and glucose were added to the ZnO NR-modified μPAD biosensor. (<b>E</b>) The graph displays representative calibration curves obtained from the ZnO NR-modified μPAD colorimetric biosensor for the detection of uric acid and glucose. Adapted with permission from Ref. [<a href="#B10-biosensors-14-00480" class="html-bibr">10</a>], copyright (2021) Elsevier B.V.</p>
Full article ">Figure 2
<p>An example of ZnO nanomaterials developed as QD fluorophores is displayed. (<b>A</b>) The panel corresponds to a transmission electron microscopy (TEM) image of ~5 nm ZnO QDs used in the detection of TC. The high-resolution image in the inset shows lattice fringes at a distance of 0.28 nm due to the inter-plane distance of the ZnO (100) plane. (<b>B</b>) The excitation and emission spectra of the ZnO QDs are shown in the contour plot. (<b>C</b>) The schematic illustrates how Eu/ZnO nanostructures were constructed for TC detection. Using the chemical reaction schemes, Eu<sup>3+</sup> was anchored on the surface of the Eu/ZnO QD fluorescent probe for the rare earth ion’s sensitization by TC. (<b>D</b>) The plot displays the emission spectra of the Eu/ZnO QDs under varying concentrations of TC up to 3 μM. The photographs in the inset were taken from the solution of the Eu/ZnO QDs with no TC (left) and 3 μM TC (right) under a 365 nm UV lamp. (<b>E</b>) The ratiometric calibration curve of the biosensor employing the fluorophores of Eu/ZnO QDs is shown as a function of TC concentration. The basis of the ratiometric analysis was I<sub>616</sub>/I<sub>530</sub>, which is the measured fluorescence intensity of Eu<sup>3+</sup> at 616 nm divided by that of ZnO QDs at 530 nm. Adapted with permission from Ref. [<a href="#B18-biosensors-14-00480" class="html-bibr">18</a>], copyright (2021) Elsevier B.V.</p>
Full article ">Figure 3
<p>Enhanced fluorescence detection of model bioanalytes facilitated by the use of ZnO NRs is exemplified. (<b>A</b>) No significant fluorescence emission was observed from fluorescein isothiocyanate-conjugated anti-immunoglobulin G antibodies (FITC–anti-IgG) when it was adsorbed on the control substrates of Si wafers, Si NRs, and stripe-patterned PMMA substrates, even with protein concentration as high as 2 mg/mL. (<b>B</b>) Markedly strong fluorescence emission was detected from 200 μg/mL FITC–anti-IgG that were deposited on ZnO NRs grown on a Si substrate either to lay flat (left panel) or to form a striped array of vertically aligned NRs (right panel). The fluorescence images in (<b>A</b>,<b>B</b>) are 25 × 25 μm<sup>2</sup> in size. The inset of the fluorescence panel in (<b>B</b>) corresponds to the top-down SEM view of the vertically grown ZnO NRs assembled into a striped array. (<b>C</b>) The plot compares the fluorescence intensity from 200 μg/mL FITC–anti-IgG molecules prepared on the different platforms of glass, Si, Si NRs, PMMA, ZnO thin film, and a striped array of ZnO NRs. (<b>D</b>) The plot of fluorescence intensity versus FITC–anti-IgG concentration compares the detection capability of ZnO NRs relative to a conventional platform of PMMA. (<b>A</b>–<b>D</b>) Adapted with permission from Ref. [<a href="#B21-biosensors-14-00480" class="html-bibr">21</a>], copyright (2006) American Chemical Society. (<b>E</b>) Protein–protein interactions were performed on a striped array of ZnO NRs. The NR platform was prepared by placing an elastomeric polymer piece of PDMS that contained two hollow chambers to simultaneously carry out protein reactions on the same platform. Protein pairs examined are fibronectin (Fn) and FITC–anti-IgG, IgG and FITC–anti-IgG, biotinylated bovine serum albumin (BBSA) and dichlorotriazinylamino fluorescein-conjugated streptavidin (DTAF–streptavidin), and IgG and DTAF–streptavidin in the reaction chambers 1 through 4. (<b>F</b>) The SEM image displays the top-down view of a striped array of ZnO NRs equally present inside all reaction chambers. As seen in the fluorescence panels, strong emissions were observable due to the reactions between interacting protein pairs in chambers 2 and 3. No detectable fluorescence signals were seen from chambers 1 and 4 due to the lack of specific protein–protein interactions. (<b>E</b>,<b>F</b>) Adapted with permission from Ref. [<a href="#B26-biosensors-14-00480" class="html-bibr">26</a>], copyright (2006) WILEY-VCH Verlag GmbH &amp; Co. KGaA, Weinheim.</p>
Full article ">Figure 4
<p>Examples of ZnO NRs used as a signal-enhancing platform for bioanalyte fluorescence detection are displayed. The bioanalytes assayed in (<b>A</b>–<b>C</b>) are the cytokines TNF-α and IL-8 in AKI patient urine samples, and those in (<b>D</b>,<b>E</b>) are the RA autoantibodies in RA patient sera. (<b>A</b>) The SEM panel displays a top-down view of vertically oriented ZnO NRs grown into a square array where the as-synthesized NRs were free from fluorescence emission, as evidenced in the inset. For the simultaneous detection of TNF-α and IL-8, sandwich-type immunoassays were performed on the ZnO NRs. The fluorescence panels of green and red were obtained from the ZnO NRs, where the two colored panels correspond to the detection channels of TNF-α and IL-8, respectively, due to their respective Alexa 488-labeled and Alexa 546-labeled secondary antibodies. (<b>B</b>) The bar graphs display fluorescence readings obtained for both biomarkers from selected patient samples. The fluorescence signals from ∼550 NR square patches per sample were analyzed to obtain the average fluorescence intensity, as well as the associated error bars reported for each patient sample. (<b>C</b>) The data show the results of interassay (data shown under *) and intra-assay (data shown under **) variability for the same patients that were carried out on three different ZnO NR arrays and five times on the same platform, respectively. (<b>A</b>–<b>C</b>) Reproduced with permission from Ref. [<a href="#B29-biosensors-14-00480" class="html-bibr">29</a>], copyright (2016) Royal Society of Chemistry. (<b>D</b>,<b>E</b>) The bar graphs summarize the diagnostic assay results of 10 RA patients and 10 healthy sera using the immunodetection scheme shown in the cartoon. The assays were performed on the platform of (<b>D</b>) ZnO NRs and (<b>E</b>) PS. The red lines inserted in the bar graphs represent the cut-off values (COVs), which were determined from the signals of the negative control samples, i.e., average plus 3 times standard deviation, measured on each platform. On the platform of ZnO NRs, all patient sera showed strong positive signals far above the COV, whereas only a couple of patient sera yielded positive signals that were barely above the COV in the assays using the conventional 96-well PS plate. The net fluorescence measured from each sample was much higher on the ZnO NRs relative to those on the PS. (<b>D</b>,<b>E</b>) Adapted with permission from Ref. [<a href="#B31-biosensors-14-00480" class="html-bibr">31</a>], copyright (2011) Elsevier B.V.</p>
Full article ">Figure 5
<p>More examples of ZnO NRs used in fluorescence-based biodetection are presented. (<b>A</b>) The schematic illustration outlines the ATP detection scheme based on the green and red fluorescence signals associated with two split aptamers—a capture aptamer intercalated with green-fluorescent SGI (cAP<sub>SGI</sub>) and a detection aptamer with an extended DNA sequence required for the formation of red fluorescent AgNCs (dAP<sub>AgNC</sub>)—while employing vertical ZnO NRs as a signal-enhancing platform. The addition of ATP resulted in a decrease in green fluorescence from cAP<sub>SGI</sub> and a multi-fold increase in red fluorescence from dAP<sub>AgNC</sub>. (<b>B</b>,<b>C</b>) The ATP concentration-dependent fluorescence intensity changes, Δf/fo, are plotted for (<b>B</b>) SGI and (<b>C</b>) AgNCs. The insets display the magnified views of the linear regimes. (<b>A</b>–<b>C</b>) Adapted with permission from Ref. [<a href="#B32-biosensors-14-00480" class="html-bibr">32</a>], copyright (2016) Elsevier B.V. (<b>D</b>) The schematic illustration depicts the glucose detection scheme of a FRET transducer made of Con A-conjugating CdSe/ZnS QDs as a donor and MG as an acceptor while employing a platform of ZnO NRs for signal enhancement. (<b>E</b>) The SEM micrographs in the top and bottom panels show a top-down view of the ZnO NRs deposited on a silicone hydrogel and a magnified view of a ZnO NR coated with CdSe/ZnS QDs. (<b>F</b>) The plot displays the linear relationship between the measured fluorescence intensity and the concentration of glucose. (<b>D</b>–<b>F</b>) Adapted with permission from Ref. [<a href="#B11-biosensors-14-00480" class="html-bibr">11</a>], copyright (2016) Elsevier B.V.</p>
Full article ">Figure 6
<p>Exemplary properties of individual ZnO NRs and their utilities in biodetection are shown. (<b>A</b>) The contour map displays the fluorescence intensity of 200 μg/mL DTAF-anti-IgG deposited on a 25 μm long ZnO NR. The DTAF-anti-IgG fluorescence signals were intense on the NR end relative to the NR side facets. The fluorescence signals on the single ZnO NR also persisted under constant irradiation, reaching a half-life (t<sub>1/2</sub>, the time point at which there is a 50% reduction of the intensity measured at t = 0) of ~1 min. They were extended much beyond those measured on the conventional substrates of PMMA and PS, whose t<sub>1/2</sub> were ~15 and 30 s, respectively. Reprinted with permission from Ref. [<a href="#B22-biosensors-14-00480" class="html-bibr">22</a>], copyright (2014) Royal Society of Chemistry. (<b>B</b>) The time-lapse fluorescence panels display the spatial and temporal emission behaviors of 1 μg/mL TRITC-anti-IgG deposited on a ZnO NR. (<b>C</b>) The FDTD simulation results in the top, middle, and bottom panels correspond to the radiation patterns from a single 576 nm emitter whose polarization is along the X, Y, and Z directions, respectively. (<b>D</b>) The nanomaterial size effect of a ZnO NR on <span class="html-italic">FINE</span> was evaluated by FDTD simulations. Far-field radiation patterns were obtained from a 517 nm electric dipole and shown for each ZnO NR of the specified length (L) and width (d). The spatial radiation patterns observed from the Z and X directions are displayed in the top and bottom simulation panels, respectively, for the ZnO NRs with the specified dimensions. (<b>E</b>) The plots show the effect of the NR length on <span class="html-italic">FINE</span> and <span class="html-italic">DoF</span>. Different concentrations of DTAF-anti-IgG were coupled to ZnO NRs of various lengths, and the DTAF-anti-IgG fluorescence signals on the ZnO NRs were measured. In all cases, the normalized fluorescence intensity value of ΔI (I<sub>avg,NRef</sub> − I<sub>avg,NRsf</sub>) indicated that the <span class="html-italic">DoF</span> increased as the NR length became longer. The subscripts of avg, NRef, and NRsf stand for average, NR end facets, and NR side facets, respectively. (<b>B</b>–<b>E</b>) Reprinted with permission from Ref. [<a href="#B23-biosensors-14-00480" class="html-bibr">23</a>], copyright (2015) Royal Society of Chemistry. (<b>F</b>) The scheme illustrates the overall fabrication process of the TNF-α sandwich immunoassay based on individual ZnO NRs. The NR immunoassay platform was integrated into a PDMS elastomer for the application of uniaxial tensile (blue) and compressive (red) strain with a microvice during fluorescence measurements. (<b>G</b>) The plots display % <span class="html-italic">FINE</span> versus % strain in NR length for TNF-α concentrations of 100, 1, and 0.1 fg/mL. Squared data points in blue correspond to the NRs undergoing tension, while circled data points in red refer to the NRs undergoing compression. The solid black lines represent the linear fits for the data points, while the black dashes show the 95% confidence ellipses. Positive values on the <span class="html-italic">x</span>-axis of % strain in NR length denote tension, while negative values indicate compression. (<b>F</b>,<b>G</b>) Reprinted with permission from Ref. [<a href="#B30-biosensors-14-00480" class="html-bibr">30</a>], copyright (2024) MDPI.</p>
Full article ">Figure 7
<p>The presented examples of biodetection exploit the PL of ZnO nanomaterials as a signal transduction means, where the changes in their emissions, such as NBE and DLE, are monitored to quantify the bioanalytes of OTA and glucose. (<b>A</b>) The schematic displays the fabrication processes for an OTA detection platform using ZnO NRs. The immunodetection layers of Protein A and anti-OTA were attained on the ZnO NRs through a series of reactions such as silanization, introduction of amino groups by treatment with APTES, introduction of aldehyde groups by modification with glutaraldehyde, covalent immobilization of Protein A, complexation with anti-OTA, and finally, BSA blocking. (<b>B</b>) The graph shows the PL spectra from steady-state conditions after the sensor platform was introduced with OTA of varying concentrations up to 20 ng/mL. (<b>C</b>) The plot displays the change in the normalized PL intensity measured at 379 nm when different concentrations of OTA were added over time. (<b>D</b>) The plot provides the ZnO NR-based immunosensor response to different OTA concentrations. The normalized sensor response plotted was determined by subtracting the normalized intensity from 1. (<b>A</b>–<b>D</b>) Adapted with permission from Ref. [<a href="#B43-biosensors-14-00480" class="html-bibr">43</a>], copyright (2017) Elsevier B.V. (<b>E</b>) The SEM micrographs show the ZnO NRs and ZnO/ZnS core/shell NRs in the left and right panels, respectively. (<b>F</b>) The diagram depicts the electron injection from the flavine moiety to the ZnO/ZnS NRs and the subsequent increase in the UV emission from the ZnO/ZnS-MAA-GOx bioconjugates. (<b>G</b>) Upon modification of the NR surfaces with MAA-GOx, the PL spectra were obtained from the platforms of ZnO NRs (blue) and ZnO/ZnS NRs (red). (<b>H</b>) The plot displays the increased PL intensity versus glucose concentration for ZnO-MAA-GOx (blue) and ZnO/ZnS-MAA-GOx (red). The glucose LODs determined for ZnO-MAA-GOx and ZnO/ZnS-MAA-GOx were 0.23 and 0.14 mM, respectively. (<b>E</b>–<b>H</b>) Adapted with permission from Ref. [<a href="#B13-biosensors-14-00480" class="html-bibr">13</a>], copyright (2011) Elsevier B.V.</p>
Full article ">Figure 8
<p>ZnO NPs coated with SiO<sub>2</sub> were used as a CL probe in HeLa cell imaging. (<b>A</b>) The graph shows the CL spectra of the ZnO NPs and ZnO/SiO<sub>2</sub> NPs in CPPO/H<sub>2</sub>O<sub>2</sub> solution. The photograph of the CPPO/H<sub>2</sub>O<sub>2</sub> solution taken in the dark after adding the ZnO/SiO<sub>2</sub> NPs is shown in the inset. (<b>B</b>) The plot displays the CL intensity versus the concentration of the ZnO/SiO<sub>2</sub> NPs. (<b>C</b>) The PL spectrum of the blank, Zn(Ac)<sub>2</sub>·2H<sub>2</sub>O, nano-ZnO NPs, ZnO NPs, and annealed ZnO NPs in the CL analysis system. The CL intensity induced in the CPPO/H<sub>2</sub>O<sub>2</sub> solution is displayed in the bar graphs for the ZnO NPs as well as other controls such as the blank solution, Zn(CH<sub>3</sub>COO)<sub>2</sub>·2H<sub>2</sub>O, nano-ZnO, and annealed ZnO NPs. The ZnO NPs and Zn(CH<sub>3</sub>COO)<sub>2</sub>·2H<sub>2</sub>O exhibit CL enhancement, whereas no impact on CL is observed from the nano-ZnO and annealed ZnO NPs. From this, the interstitial Zn atoms were deduced to be engaged in the CL process. (<b>D</b>) The schematics illustrate the CL process of the ZnO NPs. CPPO reacts with H<sub>2</sub>O<sub>2</sub>, producing 1, 2-dioxetanedione. With the injection of the ZnO NPs, a chemically initiated electron exchange luminescence (CIEEL) occurs between the intermediate and surface interstitial Zn atoms of the NPs. The interstitial Zn defects are then excited and transited to the VB, emitting blue light. The ZnO NPs exhibit yellow PL emission due to DLE associated with the electron transition from the energy level of the interstitial Zn atom to that of Zn vacancy. The CL process affects the ZnO PL, causing a red-shift in PL. (<b>E</b>) The schematic illustration represents the use of the biocompatible ZnO/SiO<sub>2</sub> NPs in CL-based cell imaging. The CL images were taken from HeLa cells cultured with the ZnO/SiO<sub>2</sub> NPs for 6 h and then added to the CPPO mixture of different H<sub>2</sub>O<sub>2</sub> concentrations, as specified in the bottom panel. (<b>F</b>) The CL intensities measured from the CL images in (<b>E</b>) are displayed. The CL intensity was directly proportional to the H<sub>2</sub>O<sub>2</sub> concentration. Adapted with permission from Ref. [<a href="#B143-biosensors-14-00480" class="html-bibr">143</a>], copyright (2020) Elsevier B.V.</p>
Full article ">Figure 9
<p>Examples of WGM and LMR biosensors fabricated from ZnO nanomaterials are shown. (<b>A</b>) The schematic illustrates a ZnO NR-based WGM biosensor and the WGM optical paths generated inside the NRs. (<b>B</b>) The plot corresponds to the PL spectrum of the Mn-doped ZnO NRs vertically oriented on a Si substrate. The PL spectrum shows the NBE emission at 378 nm, as well as multiple broad DLE peaks between 450–650 nm. WGMs of the DLEs were formed inside the optical resonators of the ZnO NRs, as annotated for each WGM peak. (<b>C</b>) The Gaussian fitting result of the WGM data in (<b>B</b>) is displayed in the plot. The WGM emission maximum occurred at 528.2 nm for the ZnO NRs, whose peak was blue-shifted to 524.4 nm upon anti-GVA immobilization to the NRs and subsequently red-shifted to 527 nm after further incubation with GVA. (<b>D</b>) The PL spectra of the ZnO NRs, as well as those treated with anti-GVA and subsequently with 1 ng/mL GVA, are presented. The inset belongs to the Gaussian-fitted WGM peaks. (<b>A</b>–<b>D</b>) Adapted with permission from Ref. [<a href="#B44-biosensors-14-00480" class="html-bibr">44</a>], copyright (2020) Elsevier B.V. (<b>E</b>) The pictorial representation shows a lossy-mode resonance (LMR) biosensor fabricated from a ZnO thin film with polypyrrole (PPY) prepared with a molecular imprinting polymer technique. The LMR sensor was subsequently used in cortisol detection. (<b>F</b>) The plot displays the absorbance wavelength of the LMR peak as a function of cortisol concentration. A blue shift of 51.23 nm in the LMR absorbance wavelength was observed for the cortisol concentration range of 10<sup>−12</sup>–10<sup>−6</sup> g/mL. (<b>G</b>) The sensitivity for the different cortisol concentrations is shown for the LMR sensor prepared with 20% ZnO/PPY. The sensitivity was evaluated from the sensor calibration curve by taking the derivative of the curve fit equation with respect to cortisol concentration. A maximum sensitivity of 12.86 nm/Log (g/mL) was yielded for 10<sup>−12</sup> g/mL of cortisol. This rate of change in the absorbance wavelength decreased with increasing cortisol concentration. (<b>E</b>–<b>G</b>) Adapted with permission from Ref. [<a href="#B48-biosensors-14-00480" class="html-bibr">48</a>], copyright (2016) Elsevier B.V.</p>
Full article ">Figure 10
<p>Exemplar uses of ZnO nanomaterials in both prism- and optical fiber-based SPR biosensors are presented. (<b>A</b>) The schematic diagram depicts a typical SPR detection setup involving a Kretschmann-type attenuated total reflection (ATR) attachment. C, H, and T indicate the flow cell, base thermistor, and hemicylinder prism, respectively. Adapted with permission from Ref. [<a href="#B173-biosensors-14-00480" class="html-bibr">173</a>], copyright (2020) American Chemical Society. (<b>B</b>) The left plot is the SPR reflectance curves obtained from the platforms of air/Au/prism, air/ZnO/Au/prism, and air/ssDNA/ZnO/Au/prism used for the detection of <span class="html-italic">N. meningitidis</span> DNA. The plot on the right is the calibration curve created from the SPR curves of the air/ssDNA/ZnO/Au/prism platform upon ssDNA hybridization with different concentrations of complementary DNA. The interaction between the probe ssDNA with increasing concentrations of target DNA resulted in a continuous shift in SPR angle towards a higher angle. The SPR biosensor exhibited a linear response range of 10–180 ng/μL <span class="html-italic">N. meningitidis</span> DNA and a LOD of 5 ng/μL. Adapted with permission from Ref. [<a href="#B50-biosensors-14-00480" class="html-bibr">50</a>], copyright (2015) Elsevier B.V. (<b>C</b>) The schematic diagram illustrates an optical fiber SPR setup used for glucose detection. An unclad portion of plastic-clad silica fiber was coated with a Ag layer onto which ZnO NRs were synthesized. The sensing probe was then immobilized with GOx. (<b>D</b>) The SEM images display the hydrothermally grown ZnO NRs in the top panel and the silica fiber surface modified with Ag/ZnO NRs/GOx in the bottom panel. (<b>E</b>) The left plot corresponds to the changes in peak absorbance wavelength as a function of glucose concentration. The shift in the peak absorbance wavelength was ~72 nm in the glucose concentration range of 0–10 mM. The plot on the right shows the sensitivity of the sensor for different glucose concentrations. The sensitivity was calculated from the change in the peak absorbance wavelength per unit change in glucose concentration. (<b>C</b>–<b>E</b>) Adapted with permission from Ref. [<a href="#B14-biosensors-14-00480" class="html-bibr">14</a>], copyright (2016) Elsevier B.V.</p>
Full article ">Figure 11
<p>(<b>A</b>) The schematic diagram depicts the fabrication process for an SERS biosensor to detect the SARS-CoV-2 spike protein. The SERS sensor platform was constructed by synthesizing vertical ZnO NRs on a Ag layer and decorating the NR side walls with Au NPs. The Au NPs on the sensor platform were then modified with MET to promote the capture of the spike protein. (<b>B</b>) The top-view SEM image displays Au NPs decorating the side walls of the ZnO NRs. Five different SERS substrates of Au NP/ZnO NRs (A1 to A5) were used in the study by varying the average diameters of the ZnO NRs: ∼155, 205, 240, 349, and 435 nm. The substrate shown in the SEM panel corresponds to A4 with an NR diameter of ~349 nm, which was reported to yield the best SERS signals. (<b>C</b>) The plots display the Raman spectrum of the SARS-CoV-2 spike protein in PBS using the SERS substrate in (<b>B</b>). The Raman signals measured at 680 cm<sup>−1</sup> were then used to create the Log–Log plot of Raman intensity versus protein concentration. Each error bar indicates the standard deviation of five different measurements from a single SERS substrate. (<b>D</b>) The detection of the SARS-CoV-2 spike protein in untreated saliva was carried out on the SERS substrate. The Raman intensity and Log–Log plots are provided for the saliva samples. Adapted with permission from Ref. [<a href="#B52-biosensors-14-00480" class="html-bibr">52</a>], copyright (2023) Elsevier B.V.</p>
Full article ">Figure 12
<p>ZnO thin film was used as an SERS biosensor for the detection of MNZ. (<b>A</b>) The field emission SEM (FESEM) image corresponds to the 6-layer ZnO thin films prepared by sol-gel dip coating and annealed at 500 °C for 30, 90, and 120 min. The last SEM micrograph is a cross-sectional view of the ZnO thin film annealed for 120 min. (<b>B</b>) The FESEM images show the Ag NPs deposited on glass substrates by the DC magnetron sputtering method for 10 and 15 s. (<b>C</b>) The schematic diagram depicts the CT processes between the MNZ molecules and the Ag NP-ZnO thin film. When excited by a 532 nm laser, electron movements can occur via multiple pathways between the various energy levels, such as the Fermi energy (E<sub>F</sub>) level of Ag NPs, the VB/CB of ZnO, the surface state energy (Ess) level of ZnO, and the HOMO/LUMO of MNZ. The Ess comes from the surface defects of ZnO. The narrowed optical band gap of the ZnO thin film after annealing can additionally support the CT. The rough and porous thin-film surface also promotes the CT by providing low reflection and high scattering of the incident light. The increased CT processes facilitate SERS through the CE mechanism. (<b>D</b>) The plots display the Raman signal of 10<sup>4</sup> ppm MNZ absorbed on the substrates of (top) the Ag NP-ZnO thin film and (middle) bare glass. The plot in the bottom panel corresponds to the blank solution of methanol (MeOH) deposited on a substrate of Ag NP-ZnO thin film. MeOH was the solvent used to prepare the MNZ solution. The Ag-ZnO substrates were prepared by 10 s sputtering of Ag NPs onto a ZnO thin film annealed for 120 min. (<b>E</b>) The plots display the SERS spectra obtained from varying MNZ concentrations absorbed on the Ag NP-ZnO thin film. The Ag/ZnO SERS substrate could detect MNZ at a concentration as low as 0.01 ppm with an SERS EF of ~10<sup>6</sup>. Adapted with permission from Ref. [<a href="#B55-biosensors-14-00480" class="html-bibr">55</a>], copyright (2023) American Chemical Society.</p>
Full article ">
13 pages, 6545 KiB  
Article
Layer-by-Layer Assembling and Capsule Formation of Polysaccharide-Based Polyelectrolytes Studied by Whispering Gallery Mode Experiments and Confocal Laser Scanning Microscopy
by Stefan Wagner, Mateusz Olszyna, Algi Domac, Thomas Heinze, Martin Gericke and Lars Dähne
Polysaccharides 2024, 5(3), 422-434; https://doi.org/10.3390/polysaccharides5030026 - 14 Aug 2024
Viewed by 648
Abstract
The layer-by-layer (LbL) assembling of oppositely charged polyelectrolytes was studied using semi-synthetic polysaccharide derivatives, namely the polycations 6-aminoethylamino-6-deoxy cellulose (ADC) and cellulose (2-(ethylamino)ethylcarbamate (CAEC), as well as the polyanion cellulose sulfate (CS). The synthetic polymers poly(allylamine) (PAH) and poly(styrene sulfonate) (PSS) were employed [...] Read more.
The layer-by-layer (LbL) assembling of oppositely charged polyelectrolytes was studied using semi-synthetic polysaccharide derivatives, namely the polycations 6-aminoethylamino-6-deoxy cellulose (ADC) and cellulose (2-(ethylamino)ethylcarbamate (CAEC), as well as the polyanion cellulose sulfate (CS). The synthetic polymers poly(allylamine) (PAH) and poly(styrene sulfonate) (PSS) were employed as well for comparison. The stepwise adsorption process was monitored by whispering gallery mode (WGM) experiments and zeta-potential measurements. Distinct differences between synthetic- and polysaccharide-based assemblies were observed in terms of the quantitative adsorption of mass and adsorption kinetics. The LbL-approach was used to prepare µm-sized capsules with the aid of porous and non-porous silica particle templates. The polysaccharide-based capsule showed a switchable permeability that was not observed for the synthetic polymer materials. At ambient pH values of 7, low-molecular dyes could penetrate the capsule wall while no permeation occurred at elevated pH values of 8. Finally, the preparation of protein-loaded LbL-capsules was studied using the combination of CAEC and CS. It was shown that high amounts of protein (streptavidin and ovomucoid) can be encapsulated and that no leaking or disintegration of the cargo macromolecules occurred during the preparation step. Based on this work, potential use in biomedical areas can be concluded, such as the encapsulation of bioactive compounds (e.g., pharmaceutical compounds, antibodies) for drug delivery or sensing purposes. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Results of whispering gallery mode adsorption experiments for the layer-by-layer assembling of bilayers of polycations (PAH: poly(allylamine); ADC: aminoethylamino-6-deoxy cellulose; CAEC: cellulose (2-(ethylamino)ethyl)carbamate) and polyanions (PSS: poly(styrene sulfonate); CS: cellulose sulfate) followed by a stability test against borate buffer (50 mM, pH value of 10) and PBS buffer displayed (<b>a</b>) over the whole process and (<b>b</b>) zoomed in for the second bilayer adsorption (after normalization to t = 0 s and WGM shift = 0 pm). Results for the quantification of the adsorption in terms of (<b>c</b>) mass of polyelectrolytes and (<b>d</b>) amount of polymer repeating units.</p>
Full article ">Figure 2
<p>Results of whispering gallery mode adsorption experiments for the layer-by-layer assembling of a bilayer of (<b>a</b>) aminoethylamino-6-deoxy cellulose/ADC and cellulose sulfate/CS or (<b>b</b>) poly(allylamine)/PAH and poly(styrene sulfonate)/PSS with two different mass concentrations for the polycation (black: 1 g/L, red: 0.1 g/L) zoomed in for the third bilayer adsorption (after normalization to t = 0 s and WGM shift = 0 pm).</p>
Full article ">Figure 3
<p>Zeta-potential (measured in pH 7 buffer) of polystyrene particles that were precoated with two double layers of poly(styrene sulfonate)/PSS and poly(allylamine)/PAH followed by alternating layers of polyanions (PSS; CS: cellulose sulfate) and polycations (PAH; ADC: aminoethylamino-6-deoxy cellulose; CAEC: cellulose (2-(ethylamino)ethyl)carbamate).</p>
Full article ">Figure 4
<p>Confocal laser scanning microscopy images capsules obtained by layer-by-layer assembling of cellulose (2-(ethylamino)ethyl)carbamate and cellulose sulfate that were functionalized at the amino groups with fluorescent 4-chloro-7-nitrobenzo-2-oxa-1,3-diazole (<b>a</b>) before and (<b>b</b>) after drying. Settings: 64× objective, 5× zoom, excitation at 405 nm, intensity of 15 %, emission from 450 to 600 nm, photomultiplier at 800 V.</p>
Full article ">Figure 5
<p>Confocal laser scanning microscopy images of capsules obtained by layer-by-layer assembling of polycations (PAH: poly(allylamine); CAEC: cellulose (2-(ethylamino)ethyl)carbamate and polyanions (PSS: poly(styrene sulfonate); CS: cellulose sulfate)) that were incubated with low-molecular sulfo-rhodamine (sRho) or high-molecular Rho-labeled PSS (PSS-Rho; 70 kDA) at different pH values. Settings: 63× objective, 1.5× zoom, excitation at 532 nm, intensity of 33 %, emission from 545 to 625 nm, photomultiplier at 650 V (<b>a</b>,<b>b</b>,<b>e</b>,<b>f</b>) or 685 V (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 6
<p>Confocal laser scanning microscopy images in transmission (<b>a</b>) and fluorescence mode (<b>b</b>) of capsules obtained by layer-by-layer assembling of cellulose (2-(ethylamino)ethyl)carbamate and cellulose sulfate after loading with rhodamine-labeled streptavidin and drying. Settings 63× objective, 1× zoom, excitation at 532 nm, intensity of 33 %, emission from 545 to 625 nm, photomultiplier at 750 V.</p>
Full article ">Figure 7
<p>Confocal laser scanning microscopy images of capsules obtained by layer-by-layer assembling of cellulose (2-(ethylamino)ethyl)carbamate and cellulose sulfate that were loaded with rhodamine-labeled (Rho) streptavidin or ovomucoid and incubated in PBS-buffer for up to 4 days. Settings: 63× objective, 1× zoom, excitation at 532 nm, intensity of 33 %, emission from 545 to 625 nm, photomultiplier at 500 V.</p>
Full article ">Scheme 1
<p>Molecular structures of polycations and polyanions used in this study.</p>
Full article ">
9 pages, 2751 KiB  
Article
Highly Sensitive Force Sensor Based on High-Q Asymmetric V-Shaped CaF2 Resonator
by Deyong Wang, Jiamin Rong, Jianglong Li, Hongbo Yue, Wenyao Liu, Enbo Xing, Jun Tang and Jun Liu
Micromachines 2024, 15(6), 751; https://doi.org/10.3390/mi15060751 - 2 Jun 2024
Viewed by 651
Abstract
Whispering gallery mode (WGM) resonators have high-quality factors and can be used in high-sensitivity sensors due to the narrow line width that allows for the detection of small external changes. In this paper, a force-sensing system based on a high-Q asymmetric V-shaped CaF [...] Read more.
Whispering gallery mode (WGM) resonators have high-quality factors and can be used in high-sensitivity sensors due to the narrow line width that allows for the detection of small external changes. In this paper, a force-sensing system based on a high-Q asymmetric V-shaped CaF2 resonator is proposed. Based on the dispersion coupling mechanism, the deformation of the resonator is achieved by loading force, and the resonant frequency is changed to determine the measurement. By adjusting the structural parameters of the asymmetric V-shaped resonator, the deformation of the resonator under force loading is improved. The experimental results show that the sensitivity of the V-shaped tip is 18.84 V/N, which determines the force-sensing resolution of 8.49 μN. This work provides a solution for force-sensing measurements based on a WGM resonator. Full article
(This article belongs to the Special Issue Recent Advances in Sensors and Sensing System Design)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Top view of the polished resonator under optical microscope. (<b>b</b>) Enlarged view of the area in which the white box is located in (<b>a</b>), with the asymmetric V-shaped structure of the area marked. (<b>c</b>) Surface morphology of polished resonators under atomic force microscopy. (<b>d</b>) Normalized transmission spectra of fabricated asymmetric V-shaped CaF<sub>2</sub> resonators.</p>
Full article ">Figure 2
<p>Experimental schematic of force sensing based on asymmetric V-shaped resonator.</p>
Full article ">Figure 3
<p>(<b>a</b>) Varying <span class="html-italic">R</span> for different forces simulated. The inset shows the deformation of <span class="html-italic">R</span> when subjected to a force of 1 N. (<b>b</b>) The variation of <span class="html-italic">n<sub>eff</sub></span> with applied force under simulation. The inset shows the resonator strain at 1 N force.</p>
Full article ">Figure 4
<p>Schematic diagram of the experimental setup for force sensing system. Isolator, ISO; fiber polarization controller, FPC; photodetector, PD; oscilloscope, OSC; force transducer, FT. Lines in red indicate optical connections; lines in black indicate circuit connections.</p>
Full article ">Figure 5
<p>(<b>a</b>) Asymmetric V-shaped profile of CaF<sub>2</sub> resonator. (<b>b</b>) The Δ<span class="html-italic">λ<sub>r</sub></span>/Δ<span class="html-italic">F</span> of the CaF<sub>2</sub> resonator as a function of ‘top’ under simulation. (<b>c</b>) The Δ<span class="html-italic">λ<sub>r</sub></span>/Δ<span class="html-italic">F</span> of the CaF<sub>2</sub> resonator as a function of ‘heights’ under simulation. Inset: the ∆<span class="html-italic">R</span>/<span class="html-italic">R</span> distributions of the resonator at bottom = 400 (700) µm. (<b>d</b>) The |Δ<span class="html-italic">λ<sub>r</sub></span>/Δ<span class="html-italic">F</span>| of the resonator as a function of ‘heights’ and ‘bottom’ for top = 50 μm under simulation. Blue sphere: structural parameters of the fabricated resonator.</p>
Full article ">Figure 6
<p>(<b>a</b>) Voltage output response of CaF<sub>2</sub> resonator under forces at two different positions. Inset: the coupling positions and the distribution of Δ<span class="html-italic">R</span>/<span class="html-italic">R</span>. (<b>b</b>) The voltage output response depends on the applied force in the case shown in (<b>a</b>). Inset: system noise.</p>
Full article ">
9 pages, 4244 KiB  
Article
Carbon Dot-Decorated Polystyrene Microspheres for Whispering-Gallery Mode Biosensing
by Anton A. Starovoytov, Evgeniia O. Soloveva, Kamilla Kurassova, Kirill V. Bogdanov, Irina A. Arefina, Natalia N. Shevchenko, Tigran A. Vartanyan, Daler R. Dadadzhanov and Nikita A. Toropov
Photonics 2024, 11(5), 480; https://doi.org/10.3390/photonics11050480 - 20 May 2024
Cited by 1 | Viewed by 1111
Abstract
Whispering gallery mode (WGM) resonators doped with fluorescent materials find impressive applications in biological sensing. They do not require special conditions for the excitation of WGM inside that provide the basis for in vivo sensing. Currently, the problem of materials for in vivo [...] Read more.
Whispering gallery mode (WGM) resonators doped with fluorescent materials find impressive applications in biological sensing. They do not require special conditions for the excitation of WGM inside that provide the basis for in vivo sensing. Currently, the problem of materials for in vivo WGM sensors are substantial since their fluorescence should have stable optical properties as well as they should be biocompatible. To address this we present WGM microresonators of 5–7 μm, where the dopant is made of carbon quantum dots (CDs). CDs are biocompatible since they are produced from carbon and demonstrate bright optical emission, which shows different bands depending on the excitation wavelength. The WGM sensors developed here were tested as label-free biosensors by detecting bovine serum albumin molecules. The results showed WGM frequency shifting, with the limit of detection down to 1016 M level. Full article
(This article belongs to the Special Issue Advancements in Optical Metamaterials)
Show Figures

Figure 1

Figure 1
<p>Spectra of water solutions: optical density (<b>a</b>) and normalized photoluminescence (<b>b</b>) of BSA (1), CDs (2), and their mixture excited at <math display="inline"><semantics> <mi>λ</mi> </semantics></math> = 488 nm (3).</p>
Full article ">Figure 2
<p>(<b>a</b>) SEM image of a polystyrene microsphere. (<b>b</b>) Normalized photoluminescence spectra of CDs solution excited at 405 nm (curve 1) and 488 nm (curve 2), CD-doped microspheres excited at 405 nm (curve 3) and 488 nm (curve 4). Inset shows a microsphere image, corresponding to spectrum 3, which was acquired by a laser scanning confocal microscope.</p>
Full article ">Figure 3
<p>(<b>a</b>) Emission spectra of polystyrene microspheres covered with carbon dots at different excitation intensities; (<b>b</b>) input-output characteristic defined as intensities of separate emission lines.</p>
Full article ">Figure 4
<p>(<b>a</b>) Emission spectra of CDs doped microspheres before and after adding BSA molecules (<math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>16</mn> </mrow> </msup> </semantics></math> M solution). WGM frequency shift extracted from emission spectra of CDs doped microspheres before and after adding BSA solution: (<b>b</b>)—<math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> </semantics></math> M, (<b>c</b>)—<math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>16</mn> </mrow> </msup> </semantics></math> M, (<b>d</b>)—<math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>18</mn> </mrow> </msup> </semantics></math> M.</p>
Full article ">Figure 5
<p>Dependence of WGM frequency shift vs. concentration of added BSA solution in logarithmic scale.</p>
Full article ">
10 pages, 3431 KiB  
Article
Underwater Low-Frequency Acoustic Wave Detection Based on a High-Q CaF2 Resonator
by Guangzheng Yuan, Jiamin Rong, Dewei Zhang, Enbo Xing, Wenyao Liu, Li Li, Jun Tang and Jun Liu
Machines 2024, 12(4), 234; https://doi.org/10.3390/machines12040234 - 3 Apr 2024
Viewed by 3571
Abstract
Whispering gallery mode (WGM) resonators with an ultra-high quality (Q) factor provide a new idea for high-precision underwater acoustic sensing. However, acoustic energy loss due to watertight encapsulation has become an urgent problem for its underwater application. In order to solve this problem, [...] Read more.
Whispering gallery mode (WGM) resonators with an ultra-high quality (Q) factor provide a new idea for high-precision underwater acoustic sensing. However, acoustic energy loss due to watertight encapsulation has become an urgent problem for its underwater application. In order to solve this problem, this paper proposes a hollowed-out array structure. The finite element simulation shows that the acoustic wave transmission loss is improved by 30 dB compared with that of the flat plate encapsulation structure. Using a calcium fluoride (CaF2) resonator with a Q factor of 1.2 × 108 as an acoustic sensitive unit, the amplitude and frequency of the loaded acoustic wave are retrieved by means of the dispersion coupling response mechanism. The resonator’s underwater experimental test range is 100 Hz–1 kHz, its acoustic sensing sensitivity level reaches −176.3 dB re 1 V/µPa @ 300 Hz, and its minimum detectable pressure can be up to 0.87 mPa/Hz1/2, which corresponds to a noise-equivalent pressure (NEP) of up to 58 dB re 1 µPa/Hz1/2. Full article
(This article belongs to the Section Machine Design and Theory)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The theoretical sensitivity of the acoustic sensor system based on the CaF<sub>2</sub> resonator. (<b>b</b>) Composition diagram of the encapsulated device.</p>
Full article ">Figure 2
<p>(<b>a</b>) <span class="html-italic">STL</span> curves of two structures obtained from simulation; the illustration shows the transmitted sound intensity at 100 Hz for both structures. (<b>b</b>) Deformation curves corresponding to different radii of CaF<sub>2</sub> at 1 Pa pressure; the illustration shows the deformation cloud of a 5 mm radius disk at 100 Hz at 1 Pa pressure.</p>
Full article ">Figure 3
<p>(<b>a</b>) The scanning transmission spectrum of the CaF<sub>2</sub> resonator has a full width at half maximum (FWHM) of 1.61 MHz, and the illustration shows the CaF<sub>2</sub> cavity. (<b>b</b>) Diagram of the packaged device.</p>
Full article ">Figure 4
<p>Diagram of the experimental setup, with the blue line representing the electrical signal path and the red line representing the optical signal path. The illustration shows the cylindrical sound barrel used in the experiment.</p>
Full article ">Figure 5
<p>Time response of standard and test hydrophones at different frequencies: (<b>a</b>) 300 Hz; (<b>b</b>) 500 Hz; (<b>c</b>) 700 Hz; (<b>d</b>) 900 Hz; the blue line represents the standard hydrophone, and the red line represents the test hydrophone.</p>
Full article ">Figure 6
<p>(<b>a</b>) Frequency response of the standard hydrophone and the hydrophone to be tested; the black curve is the hydrophone to be tested, and the red curve is the standard hydrophone. (<b>b</b>) The sensitivity curve of the hydrophone to be tested.</p>
Full article ">Figure 7
<p>(<b>a</b>) Repeatability experimental results. (<b>b</b>) Comparison between theoretical sensitivity level and experimental results.</p>
Full article ">Figure 8
<p>(<b>a</b>) CaF<sub>2</sub> resonator hydrophone with SNR at 300 Hz, and the resolution bandwidth (RBW) of 1 Hz. (<b>b</b>) MDP of CaF<sub>2</sub> resonator hydrophone.</p>
Full article ">
11 pages, 4574 KiB  
Article
Single-Mode Control and Individual Nanoparticle Detection in the Ultraviolet Region Based on Boron Nitride Microdisk with Whispering Gallery Mode
by Jiaxing Li, Qiang Li, Ransheng Chen, Qifan Zhang, Wannian Fang, Kangkang Liu, Feng Li and Feng Yun
Nanomaterials 2024, 14(6), 501; https://doi.org/10.3390/nano14060501 - 11 Mar 2024
Viewed by 1134
Abstract
Optical microcavities are known for their strongly enhanced light–matter interactions. Whispering gallery mode (WGM) microresonators have important applications in nonlinear optics, single-mode output, and biosensing. However, there are few studies on resonance modes in the ultraviolet spectrum because most materials with high absorption [...] Read more.
Optical microcavities are known for their strongly enhanced light–matter interactions. Whispering gallery mode (WGM) microresonators have important applications in nonlinear optics, single-mode output, and biosensing. However, there are few studies on resonance modes in the ultraviolet spectrum because most materials with high absorption properties are in the ultraviolet band. In this study, the performance of a microdisk cavity based on boron nitride (BN) was simulated by using the Finite-difference time-domain (FDTD) method. The WGM characteristics of a single BN microdisk with different sizes were obtained, wherein the resonance modes could be regulated from 270 nm to 350 nm; additionally, a single-mode at 301.5 nm is achieved by cascading multiple BN microdisk cavities. Moreover, we found that a BN microdisk with a diameter of 2 μm has a position-independent precise sensitivity for the nanoparticle of 140 nm. This study provides new ideas for optical microcavities to achieve single-mode management and novel coronavirus size screening, such as SARS-CoV-2, in the ultraviolet region. Full article
(This article belongs to the Special Issue Semiconductor Nanomaterials for Optoelectronic Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The WGM schematic of the BN microdisk cavity. (<b>b</b>) Refractive index (<span class="html-italic">n</span>) and extinction coefficient (<span class="html-italic">k</span>) spectra of the BN. The inset shows the UV-vis absorption spectrum of BN film with 200 nm.</p>
Full article ">Figure 2
<p>3D-FDTD simulation of the WGM characteristics of a BN microdisk with a diameter of 4 μm. (<b>a</b>) The WGM spectrum of the BN microdisk cavity. The electric field distribution (<b>b</b>) XZ plane (the solid white line indicates the microdisk edge), (<b>c</b>) XY plane, and (<b>d</b>) radial distribution at 300 nm, respectively.</p>
Full article ">Figure 3
<p>2D-FDTD simulation of the WGM resonance mode spectrum of the BN microdisk cavity with different diameters (<b>a</b>) 4 μm, (<b>b</b>) 2 μm, (<b>c</b>) 1.6 μm, and (<b>d</b>) 1 μm. The inset is the XY plane electric field distribution at the corresponding resonant wavelengths.</p>
Full article ">Figure 4
<p>(<b>a</b>) The Q factor, (<b>b</b>) the FSR and FWHM, (<b>c</b>) the azimuthal number <span class="html-italic">m</span>, and (<b>d</b>) the effective refractive index <span class="html-italic">n<sub>eff</sub></span> for the BN microdisk cavity at different sizes as a function of the wavelength.</p>
Full article ">Figure 5
<p>(<b>a</b>) Double and triple BN microdisks cascaded coupling models for single-mode regulation. (<b>c</b>) The resonant modes spectrum of different coupling types as a function of the wavelengths. (<b>b</b>,<b>d</b>,<b>e</b>) show the electric field distribution of the double microdisks cascaded coupling model at 294.3, 301.9, and 322.2 nm, respectively. (<b>f</b>) represents the field distribution of the triple at 301.5 nm.</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) are schematic diagrams of the single PS nanosphere at the antinodes and nodes (the white sphere represents detected PSNP and the black dashed line indicates the microdisk outline), respectively. (<b>c</b>,<b>d</b>) represent the resonant mode variation with different PS nanoparticle sizes at the antinodes and nodes at 309 nm. The inset shows the trend of the ΔFWHM as a function of the nanosphere’s diameter.</p>
Full article ">Figure 7
<p>The variation of FWHM (ΔFWHM) as a function of nanospheres with different sizes located at the antinodes and the nodes at 309 nm, respectively.</p>
Full article ">
15 pages, 19700 KiB  
Article
Towards a Lithium Niobate Photonic Integrated Circuit for Quantum Sensing Applications
by Jessica César-Cuello, Isabel Carnoto, Luis E. García-Muñoz and Guillermo Carpintero
Photonics 2024, 11(3), 239; https://doi.org/10.3390/photonics11030239 - 6 Mar 2024
Viewed by 1546
Abstract
Quantum transducers are key components for hybrid quantum networks, enabling the transfer of quantum states between microwave and optical photons. In the quantum community, many efforts have focused on creating and verifying the entanglement between microwave and optical fields in systems that typically [...] Read more.
Quantum transducers are key components for hybrid quantum networks, enabling the transfer of quantum states between microwave and optical photons. In the quantum community, many efforts have focused on creating and verifying the entanglement between microwave and optical fields in systems that typically operate at temperatures in the millikelvin range. Our goal is to develop an integrated microwave optical entanglement device based on a lithium niobate whispering gallery mode resonator (WGMR). To investigate the feasibility of developing such an integrated device, first, a passive photonic integrated circuit (PIC) was designed, fabricated, and characterized. The PIC was developed on a thin-film lithium niobate (TFLN) on an insulator platform, and it includes eight ring resonators and four asymmetric Mach–Zehnder interferometers. This paper presents the design and operational principles of the integrated device for microwave–optical entanglement, as well as the results of the characterization of the passive PIC. Full article
Show Figures

Figure 1

Figure 1
<p>In the scenario where the condition <math display="inline"><semantics> <mrow> <mo>(</mo> <msub> <mi>w</mi> <mrow> <mi>D</mi> <mi>F</mi> <mi>G</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>w</mi> <mrow> <mi>p</mi> <mi>u</mi> <mi>m</mi> <mi>p</mi> </mrow> </msub> <mo>−</mo> <msub> <mi>w</mi> <mrow> <mi>μ</mi> <mi>W</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> is met, the pump tone effectively generates pairs of entangled microwave and optical photons through the process of spontaneous parametric downconversion [<a href="#B16-photonics-11-00239" class="html-bibr">16</a>].</p>
Full article ">Figure 2
<p>Integrated entanglement generator and transceiver consisting of a bus waveguide, a LiNbO<sub>3</sub> ring resonator, a LiNbO<sub>3</sub> asymmetric Mach–Zehnder interferometer, two half-wavelength resonators, and two patch antennas.</p>
Full article ">Figure 3
<p>Layout of the microwave subsystem based on two half-wavelength resonators and two patch antennas.</p>
Full article ">Figure 4
<p>Radiation pattern of the patch antenna (<b>a</b>) illustrating the realized gain and (<b>b</b>) displayed with the microwave subsystem for axis reference.</p>
Full article ">Figure 5
<p>Microscope image of the photonic integrated circuit under test. Four AMZIs and eight single-coupled resonators were fabricated.</p>
Full article ">Figure 6
<p>(<b>a</b>) TFLN on insulator waveguide cross-section. (<b>b</b>) Simulated optical modes in the straight waveguide. (<b>c</b>) Simulated optical modes in the bent waveguide (50 μm radius).</p>
Full article ">Figure 7
<p>Optical response of the asymmetric structures when the light is injected into the top arm of the input directional coupler and collected from the top arm of the output directional coupler.</p>
Full article ">Figure 8
<p>Optical response of the asymmetric structures when the light is injected into the top arm of the input directional coupler and collected from the bottom arm of the output directional coupler.</p>
Full article ">Figure 9
<p>Optical response of the asymmetric structures when the light is injected into the bottom arm of the input directional coupler and collected from the top arm of the output directional coupler.</p>
Full article ">Figure 10
<p>Optical response of the asymmetric structures when the light is injected into the bottom arm of the input directional coupler and collected from the bottom arm of the output directional coupler.</p>
Full article ">Figure 11
<p>Calculated splitting ratios for the different AMZIs structures.</p>
Full article ">Figure 12
<p>Transmission spectrum of eight resonator structures.</p>
Full article ">Figure 13
<p>(<b>a</b>–<b>c</b>) Fit of the resonance dips to a Lorentzian function for structures number 6, 7, and 8, respectively.</p>
Full article ">Figure 14
<p>(<b>a</b>,<b>b</b>) Display of PDH module input signal and signal error for different parameter values: (<b>a</b>) PDH Phase = 109.3, Mod. Amp. = 2.8; (<b>b</b>) PDH Phase = 115.7, Mod. Amp. = 12.8. (<b>c</b>) Laser locked to the resonance.</p>
Full article ">
14 pages, 4766 KiB  
Article
Investigation into the Effects of Cross-Sectional Shape and Size on the Light-Extraction Efficiency of GaN-Based Blue Nanorod Light-Emitting Diode Structures
by Bohae Lee and Han-Youl Ryu
Crystals 2024, 14(3), 241; https://doi.org/10.3390/cryst14030241 - 29 Feb 2024
Viewed by 1296
Abstract
We investigated the effect of cross-sectional shape and size on the light-extraction efficiency (LEE) of GaN-based blue nanorod light-emitting diode (LED) structures using numerical simulations based on finite-difference time-domain methods. For accurate determination, the LEE and far-field pattern (FFP) were evaluated by averaging [...] Read more.
We investigated the effect of cross-sectional shape and size on the light-extraction efficiency (LEE) of GaN-based blue nanorod light-emitting diode (LED) structures using numerical simulations based on finite-difference time-domain methods. For accurate determination, the LEE and far-field pattern (FFP) were evaluated by averaging them over emission spectra, polarization, and source positions inside the nanorod. The LEE decreased as rod size increased, owing to the nanorods’ increased ratio of cross-sectional area to sidewall area. We compared circular, square, triangular, and hexagonal cross-sectional shapes in this study. To date, nanorod LEDs with circular cross sections have been mainly demonstrated experimentally. However, circular shapes were found to show the lowest LEE, which is attributed to the coupling with whispering-gallery modes. For the total emission of the nanorod, the triangular cross section exhibited the highest LEE. When the angular dependence of the LEE was calculated using the FFP simulation results, the triangular and hexagonal shapes showed relatively high LEEs for direction emission. The simulation results presented in this study are expected to be useful in designing high-efficiency nanorod LED structures with optimum nanorod shape and dimensions. Full article
(This article belongs to the Section Materials for Energy Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) A schematic side view of the finite-difference time-domain (FDTD) computational domain for the nanorod LED structure; (<b>b</b>) Cross-sectional shapes in the <span class="html-italic">x</span>-<span class="html-italic">y</span> plane for the four simulated nanorod LED structures: circle, square, equilateral triangle, and regular hexagon. The diameter of the circle and the side lengths of the equilateral triangle, square, and regular hexagon are denoted as <span class="html-italic">L</span>.</p>
Full article ">Figure 2
<p>Dipole source positions for the FDTD simulations are represented as solid dots for (<b>a</b>) circular, (<b>b</b>) square, (<b>c</b>) hexagonal, and (<b>d</b>) triangular cross sections.</p>
Full article ">Figure 3
<p>(<b>a</b>) Average LEE and (<b>b</b>) the relative difference in the average LEE are plotted as functions of the source spacing for the circular and square cross sections with <span class="html-italic">D</span> of 500 nm.</p>
Full article ">Figure 4
<p>Light-extraction efficiency (LEE) is shown as a function of the source position in the <span class="html-italic">x</span> direction for the dipole source polarized in the <span class="html-italic">x</span> and <span class="html-italic">y</span> directions. (<b>a</b>) Circular cross section with a diameter of 0.6 μm; (<b>b</b>) Square cross section with a side length of 0.6 μm.</p>
Full article ">Figure 5
<p>LEE of nanorod LEDs, which is averaged over spectrum, polarization, and source positions, is plotted as a function of <span class="html-italic">L</span> for (<b>a</b>) circular, (<b>b</b>) square, (<b>c</b>) triangular, and (<b>d</b>) hexagonal cross sections.</p>
Full article ">Figure 6
<p>The average LEE is plotted as a function of cross-sectional area (<span class="html-italic">A</span>) for the circular, triangular, square, and hexagonal cross sections.</p>
Full article ">Figure 7
<p>The electric-field intensity profiles in the QW plane of the nanorod LED structure. (<b>a</b>) Circular cross sections for <span class="html-italic">L</span> of 300, 400, 500, and 600 nm; (<b>b</b>) Triangular cross sections for <span class="html-italic">L</span> of 300, 400, 500, and 600 nm; (<b>c</b>) Hexagonal cross sections for <span class="html-italic">L</span> of 250, 300, 350, and 400 nm. The color scale bar for relative intensity is shown on the right.</p>
Full article ">Figure 8
<p>The near-field electric-field intensity profiles above the nanorod LED structure for cross sections of (<b>a</b>) circular, (<b>b</b>) square, (<b>c</b>) triangular, and (<b>d</b>) hexagonal shapes with <span class="html-italic">L</span> of 400 nm. For each cross section, the emission profile is averaged over spectrum, polarization, and source positions. The color scale bar for relative intensity is shown on the right.</p>
Full article ">Figure 9
<p>The polar plot of far-field patterns (FFPs) for different cross-sectional shapes and sizes. In this polar plot, FFP in the vertical direction corresponds to the polar angle <span class="html-italic">θ</span> of 0. (<b>a</b>) Circular cross sections for <span class="html-italic">L</span> of 200, 300, 400, and 500 nm; (<b>b</b>) Square cross sections for <span class="html-italic">L</span> of 200, 300, 400, and 600 nm; (<b>c</b>) Triangular cross sections for <span class="html-italic">L</span> of 300, 400, 600, and 800 nm; (<b>d</b>) Hexagonal cross sections for <span class="html-italic">L</span> of 100, 200, 300, and 400 nm.</p>
Full article ">Figure 10
<p>Average LEE within a collection angle (<span class="html-italic">θ<sub>c</sub></span>) for (<b>a</b>) circular, (<b>b</b>) square, (<b>c</b>) triangular, and (<b>d</b>) hexagonal cross sections.</p>
Full article ">Figure 11
<p>LEE within a collection angle (<span class="html-italic">θ<sub>c</sub></span>) as a function of cross-sectional area for the circular, triangular, square, and hexagonal nanorods. (<b>a</b>) <span class="html-italic">θ<sub>c</sub></span> = 15°; (<b>b</b>) <span class="html-italic">θ<sub>c</sub></span> = 30°.</p>
Full article ">
11 pages, 2702 KiB  
Article
Low-Threshold Anti-Stokes Raman Microlaser on Thin-Film Lithium Niobate Chip
by Jianglin Guan, Jintian Lin, Renhong Gao, Chuntao Li, Guanghui Zhao, Minghui Li, Min Wang, Lingling Qiao and Ya Cheng
Materials 2024, 17(5), 1042; https://doi.org/10.3390/ma17051042 - 24 Feb 2024
Cited by 1 | Viewed by 1347
Abstract
Raman microlasers form on-chip versatile light sources by optical pumping, enabling numerical applications ranging from telecommunications to biological detection. Stimulated Raman scattering (SRS) lasing has been demonstrated in optical microresonators, leveraging high Q factors and small mode volume to generate downconverted photons based [...] Read more.
Raman microlasers form on-chip versatile light sources by optical pumping, enabling numerical applications ranging from telecommunications to biological detection. Stimulated Raman scattering (SRS) lasing has been demonstrated in optical microresonators, leveraging high Q factors and small mode volume to generate downconverted photons based on the interaction of light with the Stokes vibrational mode. Unlike redshifted SRS, stimulated anti-Stokes Raman scattering (SARS) further involves the interplay between the pump photon and the SRS photon to generate an upconverted photon, depending on a highly efficient SRS signal as an essential prerequisite. Therefore, achieving SARS in microresonators is challenging due to the low lasing efficiencies of integrated Raman lasers caused by intrinsically low Raman gain. In this work, high-Q whispering gallery microresonators were fabricated by femtosecond laser photolithography assisted chemo-mechanical etching on thin-film lithium niobate (TFLN), which is a strong Raman-gain photonic platform. The high Q factor reached 4.42 × 106, which dramatically increased the circulating light intensity within a small volume. And a strong Stokes vibrational frequency of 264 cm−1 of lithium niobate was selectively excited, leading to a highly efficient SRS lasing signal with a conversion efficiency of 40.6%. And the threshold for SRS was only 0.33 mW, which is about half the best record previously reported on a TFLN platform. The combination of high Q factors, a small cavity size of 120 μm, and the excitation of a strong Raman mode allowed the formation of SARS lasing with only a 0.46 mW pump threshold. Full article
Show Figures

Figure 1

Figure 1
<p>A schematic illustration of the processing flow for the fabrication of the TFLN microdisk as follows: (<b>a</b>) A layer of chromium (Cr) is uniformly deposited onto the TFLN wafer (LiNbO<sub>3</sub> layer). (<b>b</b>) The femtosecond pulsed laser ablates the Cr layer into the desired microdisk pattern. (<b>c</b>) The Cr pattern is transferred to the TFLN through a chemo-mechanical polishing process, resulting in smooth and well-defined microdisk structures. (<b>d</b>) The Cr etching solution is employed to remove the Cr mask, and the silicon dioxide (SiO<sub>2</sub>) underneath the microdisk is selectively undercut in buffered HF solution, leading to the formation of suspended TFLN microdisks with the desired geometry.</p>
Full article ">Figure 2
<p>SEM images of the fabricated TFLN microdisk. (<b>a</b>) Top-view image, showing its entire structure. (<b>b</b>) Enlarged view image of the blue dashed box in (<b>a</b>), highlighting the smooth sidewall. (<b>c</b>) Side-view SEM image, exhibiting the cross-section of the microdisk. (<b>d</b>) Enlarged view of the blue dashed box in (<b>c</b>), exhibiting that the periphery of the microdisk is far from the rough pillar boundary with a distance of 18 μm.</p>
Full article ">Figure 3
<p>(<b>a</b>) Experimental setup for the generation of SRS and SARS lasing. Inset: the optical micrograph of the microdisk coupled with the tapered fiber. Here, optical spectral analyzer is denoted as OSA. (<b>b</b>) Optical microscope image of the TFLN suspended microdisk.</p>
Full article ">Figure 4
<p>(<b>a</b>) The spectrum of the generated SRS signal at 1623.4 nm wavelength. (<b>b</b>) The dependence of the Stokes Raman laser peak power on the on-chip pump power (black dots), exhibiting a threshold of only 0.33 mW and a linear growth slope of 40.6% above the threshold by linearly fitting (red line). (<b>c</b>) The spectrum of the generated SRS and SARS signals at wavelengths of 1623.4 and 1500.3 nm, respectively. (<b>d</b>) The dependence of the anti-Stokes Raman laser peak power on the on-chip pump power (black dots), showing a threshold of 0.46 mW and a linear growth slope of 1.7% above the threshold for SARS by linearly fitting (red line).</p>
Full article ">Figure 5
<p>(<b>a</b>) The spectrum at a high pump power of 2 mW. Inset: the enlarged view image of the spectrum around the SRS signal at 1623.4 nm wavelength, showing two more peaks. (<b>b</b>) Transmission spectrum around the pump mode, showing a loaded Q factor of 4.42 × 10<sup>6</sup> and a coupling efficiency of ~50%. Inset: The near infrared emission from the microdisk captured by the CCD camera with a working spectral window ranging from 400 to 1000 nm.</p>
Full article ">
17 pages, 4567 KiB  
Article
Asymmetrical Cross-Polarization Coupling in a Whispering-Gallery Microresonator
by Karleyda Sandoval and A. T. Rosenberger
Photonics 2024, 11(2), 170; https://doi.org/10.3390/photonics11020170 - 11 Feb 2024
Viewed by 1101
Abstract
Cross-polarization coupling between transverse electric (TE) and transverse magnetic (TM) whispering-gallery modes in an optical microresonator produces effects such as coupled-mode induced transparency (CMIT). The detailed analytical theory of this coupling indicates that the TE-to-TM and TM-to-TE couplings may have different strengths. Using [...] Read more.
Cross-polarization coupling between transverse electric (TE) and transverse magnetic (TM) whispering-gallery modes in an optical microresonator produces effects such as coupled-mode induced transparency (CMIT). The detailed analytical theory of this coupling indicates that the TE-to-TM and TM-to-TE couplings may have different strengths. Using an experimental setup centered around a hollow bottle resonator and polarization-sensitive throughput detection, that had been used in previous CMIT experiments, this asymmetry was confirmed and studied. By fitting the throughput spectra of both polarizations to the numerical output of a basic model, the asymmetry parameter defined as the ratio of the coupling amplitudes was determined from the output power in the polarization orthogonal to that of the input. The results of many experiments give a range for this ratio, roughly from 0.2 to 4, that agrees with the range predicted by the detailed theory. An analytical approximation of this ratio shows that the main reason for the asymmetry is a difference in the axial orders of the coupled modes. In some experimental cases, the orthogonal output is not well fitted by the model that assumes a single mode of each polarization, and we demonstrate that this fitting discrepancy can be the result of additional mode interactions. Full article
Show Figures

Figure 1

Figure 1
<p>ATS simulated in a microresonator of 90 µm radius. Upper: mode 1 throughput (blue); lower: mode 2 output (red). Parameter values: <span class="html-italic">Q</span><sub>1</sub> = 5 × 10<sup>6</sup>, <span class="html-italic">Q</span><sub>2</sub> = 2 × 10<sup>8</sup>; <span class="html-italic">M</span><sub>1</sub> = 0.9 (undercoupled), <span class="html-italic">M</span><sub>2</sub> = 0.1 (undercoupled); ∆ = 0 MHz. Assumed CPC strength and asymmetry parameter: <span class="html-italic">T<sub>c</sub></span> = 1.79 × 10<sup>–7</sup> and <span class="html-italic">b</span> = 4.</p>
Full article ">Figure 2
<p>Experimental setup [<a href="#B21-photonics-11-00170" class="html-bibr">21</a>]. Light from a tunable diode laser frequency-scanned by FG1 passes through an anamorphic prism (AP), an optical isolator (OI), an acousto-optic modulator (AOM, not used here), and wave plates (WP); a fiber coupler (FC) injects it into a fiber isolator and the polarization is further adjusted using the polarization controller (PC) before it enters the tapered section, where it is coupled into a WGM of the hollow bottle resonator (HBR). The output power undergoes polarization analysis (PA) and detection.</p>
Full article ">Figure 3
<p>Hollow bottle resonator with an equatorial (bulge) diameter of 195 μm.</p>
Full article ">Figure 4
<p>CMIT simulated in an HBR of 98 µm radius. Upper: mode 1 throughput (blue); lower: mode 2 output (red). Parameter values: <span class="html-italic">Q</span><sub>1</sub> = 5 × 10<sup>6</sup>, <span class="html-italic">Q</span><sub>2</sub> = 2 × 10<sup>8</sup>; <span class="html-italic">M</span><sub>1</sub> = 0.9 (undercoupled), <span class="html-italic">M</span><sub>2</sub> = 0.5 (undercoupled); ∆ = 0 MHz. Assumed CPC strength: <span class="html-italic">T<sub>c</sub></span> = 2.65 × 10<sup>−8</sup>. Assumed asymmetry parameter: <b>left</b>, <span class="html-italic">b</span> = 2.56; <b>right</b>, <span class="html-italic">b</span> = 6.97.</p>
Full article ">Figure 5
<p>CMIT simulated in an HBR of 90 µm radius. Upper: mode 1 throughput (blue); lower: mode 2 output (red). Black horizontal lines: values of <span class="html-italic">R</span><sub>1</sub>(0) and <span class="html-italic">R</span><sub>2</sub>(0) calculated from Equations (10) and (11). Parameter values: <span class="html-italic">Q</span><sub>1</sub> = 3.5 × 10<sup>6</sup>, <span class="html-italic">Q</span><sub>2</sub> = 4.8 × 10<sup>6</sup>; <span class="html-italic">M</span><sub>1</sub> = 0.95 (undercoupled), <span class="html-italic">M</span><sub>2</sub> = 0.38 (undercoupled); ∆ = 0 MHz. Assumed CPC strength: <span class="html-italic">T<sub>c</sub></span> = 3.11 × 10<sup>−7</sup>. Assumed asymmetry parameter: <span class="html-italic">b</span> = 1.00.</p>
Full article ">Figure 6
<p>Comparing the analytical approximation of Equation (16) for the asymmetry parameter and its numerically calculated value from the full detailed theory. The red line is a linear fit to the data points given by the equation <math display="inline"><semantics> <mrow> <msub> <mi>b</mi> <mrow> <mi>a</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>1.020</mn> <msub> <mi>b</mi> <mrow> <mi>n</mi> <mi>u</mi> </mrow> </msub> <mo>−</mo> <mn>0.024</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>ATS observed in an air-filled HBR of 172 µm radius with TE input. Upper: mode 1 throughput; lower: mode 2 output. Black: experimental spectra; blue and red: numerical fitting. Parameter values: <span class="html-italic">Q</span><sub>1</sub> = 3.13 × 10<sup>6</sup>, <span class="html-italic">Q</span><sub>2</sub> = 7.29 × 10<sup>6</sup>; <span class="html-italic">M</span><sub>1</sub> = 0.923 (undercoupled), <span class="html-italic">M</span><sub>2</sub> = 0.296 (undercoupled); ∆ = 0 MHz. Coupling strength and asymmetry parameter found by fitting: <span class="html-italic">T<sub>c</sub></span> = 3.30 × 10<sup>−7</sup>, <span class="html-italic">b</span> = 3.7. Coupling strength and asymmetry parameter found from the approximations of Equations (12) and (13): <span class="html-italic">T<sub>c</sub></span> = 3.20 × 10<sup>−7</sup>, <span class="html-italic">b</span> = 3.9.</p>
Full article ">Figure 8
<p>ATS observed in a water-filled HBR of 90 µm radius with TM input. Upper: mode 1 throughput; lower: mode 2 output. Black: experimental spectra; blue and red: numerical fitting. Parameter values: <span class="html-italic">Q</span><sub>1</sub> = 8.05 × 10<sup>6</sup>, <span class="html-italic">Q</span><sub>2</sub> = 8.05 × 10<sup>6</sup>; <span class="html-italic">M</span><sub>1</sub> = 1.0 (critically coupled), <span class="html-italic">M</span><sub>2</sub> = 0.566 (undercoupled); ∆ = 13.1 MHz. Coupling strength and asymmetry parameter found by fitting: <span class="html-italic">T<sub>c</sub></span> = 1.39 × 10<sup>−7</sup>, <span class="html-italic">b</span> = 1.0. Coupling strength and asymmetry parameter found from the approximations of Equations (12) and (13): <span class="html-italic">T<sub>c</sub></span> = 2.14 × 10<sup>−7</sup>, <span class="html-italic">b</span> = 1.2.</p>
Full article ">Figure 9
<p>ATS observed in an air-filled HBR of 98 µm radius with TE input. Upper: mode 1 throughput; lower: mode 2 output. Black: experimental spectra; blue and red: numerical fitting. Parameter values: <span class="html-italic">Q</span><sub>1</sub> = 8.8 × 10<sup>6</sup>, <span class="html-italic">Q</span><sub>2</sub> = 2.26 × 10<sup>7</sup>; <span class="html-italic">M</span><sub>1</sub> = 0.95 (undercoupled), <span class="html-italic">M</span><sub>2</sub> = 0.13 (undercoupled); ∆ = −2.3 MHz. Coupling strength and asymmetry parameter found by fitting: <span class="html-italic">T<sub>c</sub></span> = 2.26 × 10<sup>−8</sup>, <span class="html-italic">b</span> = 3.7. Coupling strength and asymmetry parameter found from the approximations of Equations (12) and (13): <span class="html-italic">T<sub>c</sub></span> = 2.16 × 10<sup>−8</sup>, <span class="html-italic">b</span> = 2.7.</p>
Full article ">Figure 10
<p>ATS observed in an air-filled HBR of 89 µm radius with TM input. Upper: mode 1 throughput; lower: mode 2 output. Black: experimental spectra; blue and red: numerical fitting. Parameter values: <span class="html-italic">Q</span><sub>1</sub> = 7.06 × 10<sup>6</sup>, <span class="html-italic">Q</span><sub>2</sub> = 9.39 × 10<sup>6</sup>; <span class="html-italic">M</span><sub>1</sub> = 1.0 (critically coupled), <span class="html-italic">M</span><sub>2</sub> = 0.30 (overcoupled); ∆ = −21 MHz. Coupling strength and asymmetry parameter found by fitting: <span class="html-italic">T<sub>c</sub></span> = 2.97 × 10<sup>−7</sup>, <span class="html-italic">b</span> = 0.62. Coupling strength and asymmetry parameter found from the approximations of Equations (12) and (13): <span class="html-italic">T<sub>c</sub></span> = 5.87 × 10<sup>−7</sup>, <span class="html-italic">b</span> = 0.48.</p>
Full article ">Figure 11
<p>ATS observed in an air-filled HBR of 172 µm radius with TE input. Upper: mode 1 throughput; lower: mode 2 output. Black: experimental spectra; blue and red: numerical fitting. Parameter values: <span class="html-italic">Q</span><sub>1</sub> = 3.65 × 10<sup>6</sup>, <span class="html-italic">Q</span><sub>2</sub> = 1.37 × 10<sup>7</sup>; <span class="html-italic">M</span><sub>1</sub> = 0.933 (undercoupled), <span class="html-italic">M</span><sub>2</sub> = 0.938 (overcoupled); ∆ = −6.0 MHz. Coupling strength and asymmetry parameter found by fitting: <span class="html-italic">T<sub>c</sub></span> = 3.39 × 10<sup>−7</sup>, <span class="html-italic">b</span> = 2.3. Coupling strength and asymmetry parameter found from the approximations of Equations (12) and (13): <span class="html-italic">T<sub>c</sub></span> = 3.12 × 10<sup>−7</sup>, <span class="html-italic">b</span> = 1.4.</p>
Full article ">Figure 12
<p>ATS observed in an air-filled HBR of 172 µm radius with TE input. Upper: mode 1 throughput; lower: mode 2 output. Black: experimental spectra; blue and red: numerical fitting. Parameter values: <span class="html-italic">Q</span><sub>1</sub> = 9.99 × 10<sup>6</sup>, <span class="html-italic">Q</span><sub>2</sub> = 1.50 × 10<sup>7</sup>; <span class="html-italic">M</span><sub>1</sub> = 1.0 (critically coupled), <span class="html-italic">M</span><sub>2</sub> = 1.0 (critically coupled); ∆ = −2.77 MHz. Coupling strength and asymmetry parameter found by fitting: <span class="html-italic">T<sub>c</sub></span> = 4.09 × 10<sup>−8</sup>, <span class="html-italic">b</span> = 0.62. Coupling strength and asymmetry parameter found from the approximations of Equations (12) and (13): <span class="html-italic">T<sub>c</sub></span> = 4.00 × 10<sup>−8</sup>, <span class="html-italic">b</span> = 0.66.</p>
Full article ">Figure 13
<p>Simulation involving three modes nearly coresonant with each other, one of polarization 1 and two of polarization 2. Blue: polarization 1 throughput, showing ATS due to the interaction of the two orthogonal modes coupled by CPC. Red: polarization 2 output, showing additional sharp structure due to the interaction of the two modes of polarization 2.</p>
Full article ">
10 pages, 1947 KiB  
Article
Active Optical Tuning of Azopolymeric Whispering Gallery Mode Microresonators for Filter Applications
by Gabriel H. A. Jorge, Filipe A. Couto, Juliana M. P. Almeida, Victor A. S. Marques, Marcelo B. Andrade and Cleber R. Mendonça
Photonics 2024, 11(2), 167; https://doi.org/10.3390/photonics11020167 - 9 Feb 2024
Viewed by 1317
Abstract
Light confinement provided by whispering gallery mode (WGM) microresonators is especially useful for integrated photonic circuits. In particular, the tunability of such devices has gained increased attention for active filtering and lasering applications. Traditional lithographic approaches for fabricating such devices, especially Si-based ones, [...] Read more.
Light confinement provided by whispering gallery mode (WGM) microresonators is especially useful for integrated photonic circuits. In particular, the tunability of such devices has gained increased attention for active filtering and lasering applications. Traditional lithographic approaches for fabricating such devices, especially Si-based ones, often restrict the device’s tuning due to the material’s inherent properties. Two-photon polymerization (2PP) has emerged as an alternative fabrication technique of sub-diffraction resolution 3D structures, in which compounds can be incorporated to further expand their applications, such as enabling active devices. Here, we exploited the advantageous characteristics of polymer-based devices and produced, via 2PP, acrylic-based WGM hollow microcylinders incorporated with the azoaromatic chromophore Disperse Red 13 (DR13). Within telecommunication range, we demonstrated the tuning of the microresonator’s modes by external irradiation within the dye’s absorption peak (at 514 nm), actively inducing a blueshift at a rate of 1.2 nm/(Wcm−2). Its thermo-optical properties were also investigated through direct heating, and the compatibility of both natural phenomena was also confirmed by finite element simulations. Such results further expand the applicability of polymeric microresonators in optical and photonic devices since optically active filtering was exhibited. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Illustration of the evanescent coupling of the resonator using a tapered optical fiber. (<b>b</b>) External excitation of the WGM resonator using an Ar-ion laser at 514 nm. (<b>c</b>) Schematic representation of heating the microresonator using a ceramic resistor.</p>
Full article ">Figure 2
<p>Absorption spectra of a 15 µm thick polymerized film doped with DR13. The inset displays a SEM micrography of a microresonator sharing the same composition, fabricated by direct femtosecond laser writing via 2PP.</p>
Full article ">Figure 3
<p>(<b>a</b>) Transmission spectra of a 2 µm waist taper coupled to a DR13-doped polymeric microresonator for seven minutes. The free spectral range values of ~9.3 nm are displayed. (<b>b</b>) Zoom-in at the deepest resonance (shaded in part a) along with the Q-factor and full width half maximum obtained by fitting.</p>
Full article ">Figure 4
<p>Transmission spectra of the DR13-doped polymeric microresonator for increasing (<b>a</b>) and decreasing (<b>b</b>) external irradiances (Ar-ion laser at 514 nm). (<b>c</b>) Resonance (1515.8 nm) shift as a function of the external laser irradiance increase (blue) and decrease (red). The blue and red arrows are guides to the eye, respectively, for increasing and decreasing irradiation.</p>
Full article ">Figure 5
<p>Resonance (1519.47 nm) shift as a function of the temperature increase (blue) and decrease (red). The blue and red arrows are guides to the eye, respectively, for increasing and decreasing temperature.</p>
Full article ">Figure 6
<p>(<b>a</b>) Geometry of a simulated microresonator along with the region in which its heat source from irradiation is defined; (<b>b</b>) temperature map of the resonator submitted to an arbitrary irradiance after transient time; and (<b>c</b>) radial displacement map of the same resonator.</p>
Full article ">
12 pages, 3580 KiB  
Article
The Growth Mechanism, Luminescence, and Lasing of Polyhedral ZnO Microcrystals with Whispering-Gallery Modes
by Ludmila A. Zadorozhnaya, Andrey P. Tarasov and Vladimir M. Kanevsky
Photonics 2023, 10(12), 1328; https://doi.org/10.3390/photonics10121328 - 29 Nov 2023
Cited by 2 | Viewed by 1183
Abstract
This work studies the features of the formation of isometric polyhedral ZnO microcrystals that provide stimulated emission and whispering-gallery-mode (WGM) lasing in the near-UV range. For this purpose, the growth stages of such crystals in the process of gas-transport synthesis and the luminescent [...] Read more.
This work studies the features of the formation of isometric polyhedral ZnO microcrystals that provide stimulated emission and whispering-gallery-mode (WGM) lasing in the near-UV range. For this purpose, the growth stages of such crystals in the process of gas-transport synthesis and the luminescent properties of the structures obtained at each stage were investigated. It was shown that the growth of laser microcrystals begins with the formation of microspheroids with thin ZnO shells. Such spheroids exhibit mainly white luminescence with a small contribution of near-UV emission. Increasing the synthesis duration results in thickening and faceting of the spheroid shells, as well as a decrease in the contribution of the yellow–red component to the luminescence spectrum. At the same time, ZnO microcrystallites nucleate and grow inside the spheroids, using as a material the remains of a liquid zinc drop and oxygen entering the spheroids through their shells. Such growth conditions allow them to take on an equilibrium polyhedral shape. Eventually, upon destruction of the spheroid shell, a polyhedral ZnO microcrystal supporting WGMs is observed. Full article
(This article belongs to the Special Issue Women’s Special Issue Series: Photonics)
Show Figures

Figure 1

Figure 1
<p>SEM-images of samples S1 (<b>a</b>), S2 (<b>b</b>), and S3 (<b>c</b>).</p>
Full article ">Figure 2
<p>XRD patterns of samples S1 (<b>top</b>) and S3 (<b>bottom</b>).</p>
Full article ">Figure 3
<p>PL spectra of samples registered in the near-UV and visible ranges under low-intensity <span class="html-italic">cw</span> excitation (<b>a</b>,<b>b</b>) and pulsed laser excitation with power density, <span class="html-italic">ρ</span><sub>exc</sub> ≈ 15 kW/cm<sup>2</sup> (<b>c</b>). The spectra are normalized to the NBEL band maximum. Figures (<b>b</b>,<b>c</b>) show the Gaussian decomposition of the visible part of the PL spectra (DL) plotted vs. photon energy.</p>
Full article ">Figure 4
<p>The examples of NBEL spectrum evolution with increasing excitation power density, <span class="html-italic">ρ</span><sub>exc</sub>: (<b>a</b>) spontaneous emission of sample S1; (<b>b</b>) spontaneous emission of sample S2; (<b>c</b>) ASE in sample S3; (<b>d</b>) lasing in sample S3.</p>
Full article ">
12 pages, 4592 KiB  
Article
Whispering-Gallery Mode Micro-Ring Resonator Integrated with a Single-Core Fiber Tip for Refractive Index Sensing
by Monika Halendy and Sławomir Ertman
Sensors 2023, 23(23), 9424; https://doi.org/10.3390/s23239424 - 26 Nov 2023
Cited by 7 | Viewed by 1603
Abstract
A micro-ring resonator structure was fabricated via the two-photon polymerization technique directly on a single-mode fiber tip and tested for refractive index sensing application. The micro-ring structure was used to excite whispering-gallery modes, and observations of the changes in the resonance spectrum introduced [...] Read more.
A micro-ring resonator structure was fabricated via the two-photon polymerization technique directly on a single-mode fiber tip and tested for refractive index sensing application. The micro-ring structure was used to excite whispering-gallery modes, and observations of the changes in the resonance spectrum introduced by changes in the refractive index of the environment served as the sensing principle. The proposed structure has the advantages of a very simple design, allowing for measurements in reflection mode, relatively easy and fast fabrication and integration with a single tip of a standard single-mode fiber, which allowed for quick and convenient measurements in the optical setup. The performance of the structure was characterized, and the resonant spectrum giving high potential for refractive index sensing was measured. Future perspectives of the research are addressed. Full article
(This article belongs to the Special Issue Developments and Applications of Optical Fiber Sensors)
Show Figures

Figure 1

Figure 1
<p>A schematic view of the proposed structure, illustrating light propagation inside it—scenario concerning only one direction. The red circular arrow indicates the “ring” path (mentioned later in the paper), and the yellow and blue arrows indicate the “loop” path.</p>
Full article ">Figure 2
<p>Microscope images of the printed structures with the ring diameter equal to (<b>a</b>) 40 μm, (<b>b</b>) 34 μm, (<b>c</b>) 28 μm.</p>
Full article ">Figure 3
<p>The measurement setup.</p>
Full article ">Figure 4
<p>Measured spectra for the printed structures with the ring diameter equal to (<b>a</b>) 40 μm, (<b>b</b>) 34 μm, (<b>c</b>) 28 μm. Insets show the region of the best extinction ratio.</p>
Full article ">Figure 4 Cont.
<p>Measured spectra for the printed structures with the ring diameter equal to (<b>a</b>) 40 μm, (<b>b</b>) 34 μm, (<b>c</b>) 28 μm. Insets show the region of the best extinction ratio.</p>
Full article ">Figure 5
<p>Lumerical “Movie” monitor screenshots of (<b>a</b>) light propagation during the first roundtrip, (<b>b</b>) light propagation during the second roundtrip, (<b>c</b>) light propagation during the second roundtrip after relocating waveguide arms.</p>
Full article ">Figure 6
<p>A comparison between the output resonance spectra provided by Lumerical “Time” monitor before and after aforementioned adjustment.</p>
Full article ">Figure 7
<p>The measured spectrum of the simulated structure.</p>
Full article ">Figure 8
<p>(<b>a</b>) A microscope image and (<b>b</b>) the measured spectrum for the thickened micro-resonator structure surrounded by protective enclosure.</p>
Full article ">Figure 9
<p>(<b>a</b>) Measured spectra for varying ambient RI, (<b>b</b>) shifts of a resonant dip for varying ambient RI.</p>
Full article ">
6 pages, 1620 KiB  
Brief Report
Automatic Alignment Method for Controlled Free-Space Excitation of Whispering-Gallery Resonances
by Davide D’Ambrosio, Marialuisa Capezzuto, Antonio Giorgini, Pietro Malara, Saverio Avino and Gianluca Gagliardi
Sensors 2023, 23(21), 9007; https://doi.org/10.3390/s23219007 - 6 Nov 2023
Cited by 1 | Viewed by 910
Abstract
Whispering-gallery mode microresonators have gained wide popularity as experimental platforms for different applications, ranging from biosensing to nonlinear optics. Typically, the resonant modes of dielectric microresonators are stimulated via evanescent wave coupling, facilitated using tapered optical fibers or coupling prisms. However, this method [...] Read more.
Whispering-gallery mode microresonators have gained wide popularity as experimental platforms for different applications, ranging from biosensing to nonlinear optics. Typically, the resonant modes of dielectric microresonators are stimulated via evanescent wave coupling, facilitated using tapered optical fibers or coupling prisms. However, this method poses serious shortcomings due to fabrication and access-related limitations, which could be elegantly overcome by implementing a free-space coupling approach; although additional alignment procedures are needed in this case. To address this issue, we have developed a new algorithm to excite the microresonator automatically. Here, we show the working mechanism and the preliminary results of our experimental method applied to a home-made silica microsphere, using a visible laser beam with a spatial light modulator and a software control. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental setup. A grating-like phase mask is imaged onto the SLM panel to tilt the laser beam by acting on the reflection angle r (shown in turquoise). Light scattered from the resonator is imaged using a CCD camera through a 1:1 telescope. The telescope, and a second photodiode (PD) for backscattering detection, are omitted for sake of clarity.</p>
Full article ">Figure 2
<p>Images recorded with the CCD camera when a WGM is excited (<b>a</b>) or not (<b>b</b>). On the corresponding lower panels (<b>c</b>,<b>d</b>), we report a 1 D integration of the total number of counts. We ruled out via software the contribution of the two stray light spots, visible in both upper panels, so as to obtain a robust selection criterion to distinguish among the system’s available alignment conditions.</p>
Full article ">Figure 3
<p>Microresonator’s self-aligned modes, normalized for ease of visualization. In each plot, the tilt angle and the number of iterations are also reported. For N = 100 iterations of the algorithm, the system tends to align on the mode in the lower panel, while for lower N values, the system finds local maxima of the scattered light that correspond to lower-Q whispering-gallery modes (from the top, panels 1 and 2).</p>
Full article ">Figure 4
<p>Microresonator’s self-aligned modes in the presence of dust particles (normalized). Split modes are visible in the spectra, revealing the presence of scatterers on the surface of the WGMR.</p>
Full article ">
17 pages, 3033 KiB  
Article
Confirmation of Dissipative Sensing Enhancement in a Microresonator Using Multimode Input
by Sreekul Raj Rajagopal, Limu Ke, Karleyda Sandoval and Albert T. Rosenberger
Sensors 2023, 23(21), 8700; https://doi.org/10.3390/s23218700 - 25 Oct 2023
Viewed by 748
Abstract
Optical microresonators have proven to be especially useful for sensing applications. In most cases, the sensing mechanism is dispersive, where the resonance frequency of a mode shifts in response to a change in the ambient index of refraction. It is also possible to [...] Read more.
Optical microresonators have proven to be especially useful for sensing applications. In most cases, the sensing mechanism is dispersive, where the resonance frequency of a mode shifts in response to a change in the ambient index of refraction. It is also possible to conduct dissipative sensing, in which absorption by an analyte causes measurable changes in the mode linewidth and in the throughput dip depth. If the mode is overcoupled, the dip depth response can be more sensitive than the linewidth response, but overcoupling is not always easy to achieve. We have recently shown theoretically that using multimode input to the microresonator can enhance the dip-depth sensitivity by a factor of several thousand relative to that of single-mode input and by a factor of nearly 100 compared to the linewidth sensitivity. Here, we experimentally confirm these enhancements using an absorbing dye dissolved in methanol inside a hollow bottle resonator. We review the theory, describe the setup and procedure, detail the fabrication and characterization of an asymmetrically tapered fiber to produce multimode input, and present sensing enhancement results that agree with all the predictions of the theory. Full article
(This article belongs to the Special Issue Feature Papers in Optical Sensors 2023)
Show Figures

Figure 1

Figure 1
<p>Illustration of the experimental setup.</p>
Full article ">Figure 2
<p>Closeup of the coupling region with the HBR microresonator and the tapered fiber (highlighted in red). In this photograph, the fiber and HBR have been separated for clarity. The reservoirs for the solvent and analyte are also shown.</p>
Full article ">Figure 3
<p>Asymmetric tapered fiber with nonadiabatic downtaper and adiabatic uptaper. Light propagates from left to right.</p>
Full article ">Figure 4
<p>Delineation curves and plot of the cladding taper angle Ω as a function of the inverse taper ratio for an adiabatic taper to a waist radius of 1.15 μm.</p>
Full article ">Figure 5
<p>Delineation curves and plots of cladding taper angles Ω as a function of the inverse taper ratio for an asymmetric tapered fiber with a waist radius <span class="html-italic">r<sub>w</sub></span> = 1.16 µm.</p>
Full article ">Figure 6
<p>Throughput WGM spectra with input from tapered fibers of the same waist radius, <span class="html-italic">r<sub>w</sub></span> = 1.16 µm. (<b>a</b>) Asymmetic tapered fiber. Upward displacement indicates an increasing analyte concentration (the order of traces is same as that in the legend); the bottom trace is for methanol only, with input modes out of phase. (<b>b</b>) Symmetric tapered fiber. Oscilloscope screenshot showing the mode of interest in the yellow box (the same mode as highlighted in (a); methanol only), very close to being critically coupled, as the bottom of the throughput dip is nearly at zero (the red line).</p>
Full article ">Figure 7
<p>Fractional change in the dip depth plotted as a function of the analyte concentration, with the asymmetric tapered fiber used to couple light into and out of the microresonator.</p>
Full article ">Figure 8
<p>Fractional change in the dip depth plotted as a function of the analyte concentration, with the symmetric tapered fiber used to couple light into and out of the microresonator.</p>
Full article ">Figure 9
<p>Fractional change in the linewidth plotted as a function of the analyte concentration, with the asymmetric tapered fiber used to couple light into and out of the microresonator.</p>
Full article ">Figure 10
<p>Fractional change in the linewidth plotted as a function of the analyte concentration, with the symmetric tapered fiber used to couple light into and out of the microresonator.</p>
Full article ">
Back to TopTop