[go: up one dir, main page]

Next Issue
Volume 6, June
Previous Issue
Volume 5, December
 
 

Photonics, Volume 6, Issue 1 (March 2019) – 34 articles

Cover Story (view full-size image): Multi-spectral midwave-infrared (mid-IR) lasers are demonstrated by directly bonding quantum cascade epitaxial gain layers to silicon-on-insulator (SOI) waveguides with arrayed waveguide grating (AWG) multiplexers. Arrays of distributed feedback and distributed Bragg-reflection quantum cascade lasers (QCLs) emitting at 4.7 µm wavelength are coupled to AWGs on the same chip. Low-loss spectral beam combining allows for brightness scaling by coupling the light generated by multiple input QCLs into the fundamental mode of a single output waveguide. Promising results are demonstrated and further improvements are in progress. This device may lead to compact and sensitive chemical detection systems using absorption spectroscopy across a broad spectral range in the mid-IR, as well as a high-brightness multi-spectral source for power scaling. View Paper here.
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
14 pages, 4366 KiB  
Article
Quantitative Analysis of 4 × 4 Mueller Matrix Transformation Parameters for Biomedical Imaging
by Wei Sheng, Weipeng Li, Ji Qi, Teng Liu, Honghui He, Yang Dong, Shaoxiong Liu, Jian Wu, Daniel S. Elson and Hui Ma
Photonics 2019, 6(1), 34; https://doi.org/10.3390/photonics6010034 - 26 Mar 2019
Cited by 32 | Viewed by 4831
Abstract
Mueller matrix polarimetry is a potentially powerful technique for obtaining microstructural information of biomedical specimens. Thus, it has found increasing application in both backscattering imaging of bulk tissue samples and transmission microscopic imaging of thin tissue slices. Recently, we proposed a technique to [...] Read more.
Mueller matrix polarimetry is a potentially powerful technique for obtaining microstructural information of biomedical specimens. Thus, it has found increasing application in both backscattering imaging of bulk tissue samples and transmission microscopic imaging of thin tissue slices. Recently, we proposed a technique to transform the 4 × 4 Mueller matrix elements into a group of parameters, which have explicit associations with specific microstructural features of samples. In this paper, we thoroughly analyze the relationships between the Mueller matrix transformation parameters and the characteristic microstructures of tissues by using experimental phantoms and Monte Carlo simulations based on different tissue mimicking models. We also adopt quantitative evaluation indicators to compare the Mueller matrix transformation parameters with the Mueller matrix polar decomposition parameters. The preliminary imaging results of bulk porcine colon tissues and thin human pathological tissue slices demonstrate the potential of Mueller matrix transformation parameters as biomedical diagnostic indicators. Also, this study provides quantitative criteria for parameter selection in biomedical Mueller matrix imaging. Full article
(This article belongs to the Special Issue Biomedical Photonics Advances)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic of the Mueller matrix configuration for both backscattering and transmission measurements. L1, L2, L3: lens; P1, P2, P3: polarizer; R1, R2, R3: quarter-wave plate. (<b>b</b>) Silk phantom; the diameter of the silk fiber is 1.5 μm and its refractive index is 1.56; the scattering coefficient of the silk layer is 70 cm<sup>−1</sup>, the outer diameter of the phantom is 1.5 cm and its thickness is 2 mm.</p>
Full article ">Figure 2
<p>(<b>a</b>) A 2D backscattering Mueller matrix image of silk phantom. The color code for m12, m13, m21, and m31 is from −0.1 to 0.1. The color code for m23 and m32 is from −0.2 to 0.2. The color code for other elements is from −1 to 1. (<b>b</b>) A 2D transmission Mueller matrix image of birefringent gradient-index (GRIN) lens. The color code for all 16 elements is from −1 to 1. Azimuthal dependent curves of the Mueller matrix elements for (<b>c</b>) the silk phantom and (<b>d</b>) the GRIN lens. All the Mueller matrix elements are normalized by m11.</p>
Full article ">Figure 3
<p>Mueller matrix derived parameters of the samples: (<b>a</b>) Mueller matrix polar decomposition (MMPD) and Mueller matrix transformation (MMT) parameters of the silk phantom. (<b>b</b>) MMPD and MMT parameters of the GRIN lens. Since the m14 values for both samples are close to 0 (<a href="#photonics-06-00034-f002" class="html-fig">Figure 2</a>a–d), the images of <span class="html-italic">D</span> (not shown) and <span class="html-italic">t</span><sub>2</sub> are almost the same.</p>
Full article ">Figure 4
<p>Monte Carlo (MC) simulation results of bi-component sphere model in the backward direction: (<b>a</b>) MMPD parameter Δ. (<b>b</b>) MMT parameter <span class="html-italic">b</span>. (<b>c</b>) <span class="html-italic">T</span> of Δ and <span class="html-italic">b</span>; the codomains of both parameters are [0,1]. (<b>d</b>) The <span class="html-italic">r</span><sup>2</sup> of Δ and <span class="html-italic">b</span>.</p>
Full article ">Figure 5
<p>MC simulation results of sphere-cylinder model in backward direction: (<b>a</b>) MMPD parameters <span class="html-italic">D</span> and <span class="html-italic">δ</span>. (<b>b</b>) MMT parameters <span class="html-italic">t</span><sub>2</sub> and <span class="html-italic">t</span><sub>3</sub>. (<b>c</b>) <span class="html-italic">T</span> of <span class="html-italic">D</span> and <span class="html-italic">t</span><sub>2</sub>; the codomains of <span class="html-italic">D</span> and <span class="html-italic">t</span><sub>2</sub> are [0,1]. (<b>d</b>) The <span class="html-italic">r</span><sup>2</sup> of <span class="html-italic">D</span> and <span class="html-italic">t</span><sub>2</sub>.</p>
Full article ">Figure 6
<p>MC simulation results of a sphere-birefringence model in both forward and backward scattering directions: (<b>a</b>) MMPD parameters <span class="html-italic">D</span> and <span class="html-italic">δ</span>. (<b>b</b>) MMT parameter <span class="html-italic">t</span><sub>2</sub> and <span class="html-italic">t</span><sub>3</sub>. (<b>c</b>) <span class="html-italic">T</span> of <span class="html-italic">δ</span> and <span class="html-italic">t</span><sub>3</sub>; the codomain of <span class="html-italic">δ</span> is [0, π]; the codomain of <span class="html-italic">t</span><sub>3</sub> is [0,1]. (<b>d</b>) The <span class="html-italic">r</span><sup>2</sup> of <span class="html-italic">δ</span> and <span class="html-italic">t</span><sub>3</sub>.</p>
Full article ">Figure 7
<p>Mueller matrix derived parameters of: (<b>a</b>) ex vivo porcine colon tissue; all the color scales are set to 0.3 times the codomains of the parameters. It should be noticed that there are areas with abnormally small values of Δ and large values of <span class="html-italic">b</span>, which result from the specular reflection induced pixel saturation. (<b>b</b>) Human breast invasive ductal carcinoma tissue: 12 μm-thick, unstained, and dewaxed; all the color scales are 0.2 times the codomains of the parameters.</p>
Full article ">Figure 8
<p>Comparison between T<span class="html-italic"><sub>δ</sub></span>: T values of the MMPD parameter <span class="html-italic">δ,</span> and T<span class="html-italic"><sub>t</sub></span><sub>3</sub>: T values of the MMT parameter <span class="html-italic">t</span><sub>3</sub> for 90 human pathological tissue slices. X-axis: T<span class="html-italic"><sub>δ</sub>.</span> Y-axis: T<span class="html-italic"><sub>t</sub></span><sub>3</sub>. Red circles: Crohn-Crohn disease tissues. Blue asterisks: intestine-intestinal tuberculosis tissues. Green triangles: breast-breast carcinoma tissues.</p>
Full article ">
12 pages, 5224 KiB  
Article
Verification of Non-thermal Effects of 0.3–0.6 THz-Waves on Human Cultured Cells
by Noriko Yaekashiwa, Hisa Yoshida, Sato Otsuki, Shin’ichiro Hayashi and Kodo Kawase
Photonics 2019, 6(1), 33; https://doi.org/10.3390/photonics6010033 - 25 Mar 2019
Cited by 10 | Viewed by 4431
Abstract
Recent progress has been made in the development of terahertz (THz) waves for practical applications. Few studies that have assessed the biological effects of THz waves have been reported, and the data currently available regarding the safety of THz waves is inadequate. In [...] Read more.
Recent progress has been made in the development of terahertz (THz) waves for practical applications. Few studies that have assessed the biological effects of THz waves have been reported, and the data currently available regarding the safety of THz waves is inadequate. In this study, the effect of THz wave exposure on two cultured cells was assessed using a widely tunable THz source with a 0.3–0.6 THz frequency range, which can be used and increased in one GHz increments. The THz waves applied to the cultured cells were weak enough such that any thermal effects could be disregarded. The influence of THz wave exposure on both the proliferative and metabolic activities of these cells was investigated, as well as the extent of the thermal stress placed on the cells. In this work, no measurable effect on the proliferative or metabolic activities of either cell type was observed following the exposure to THz waves. No differences in the quantity of cDNA related to heat shock protein 70 was detected in either the sham or exposure group. As such, no differences in cellular activity between cells exposed to THz waves and those not exposed were observed. Full article
(This article belongs to the Special Issue Terahertz Photonics)
Show Figures

Figure 1

Figure 1
<p>Scheme of the exposure set-up showing the terahertz (THz) wave source in the incubator chamber for measuring cell proliferation (<b>a</b>) and cell viability (<b>b</b>).</p>
Full article ">Figure 2
<p>The frequency dependence on the irradiation intensity inside the culture well (<b>a</b>) and 96-well plate (<b>b</b>). (<b>a</b>) The irradiated THz wave covered the well through an aluminum waveguide. The intensity was less than 1 μW/cm<sup>2</sup> including coupling and propagating loss by the waveguide and an absorption loss by the bottom of the cell. (<b>b</b>) The irradiated THz wave covered all of the 96-well plate through free space with the intensity approximately 0.1 μW/cm<sup>2</sup>, including an absorption loss by the bottom of the well.</p>
Full article ">Figure 3
<p>Uni-Travelling-Carrier Photodiode (UTC-PD) optical system on the inverted microscope.</p>
Full article ">Figure 4
<p>Growth curves of cells that have been exposed to 0.3–0.6 THz waves. The solid line represents exposed cells, the broken line represents unexposed cells, the dash line represents adding 20% of DMSO and the gray line represents the culture medium without cells. Cells were exposed to waves from 0.3 THz to 0.6 THz, increasing at a rate of 1.0 GHz every 18 min and 34 s, over a total of 94 h. (<b>a</b>) NB1RGB cell proliferation curve. (<b>b</b>) HCE-T cell proliferation curve. Data were independently collected three times and the data measured at each frequency was averaged.</p>
Full article ">Figure 5
<p>Cells following exposure to 0.3–0.6 THz waves for 94 h (x5).</p>
Full article ">Figure 6
<p>Viability of cells exposed to frequencies between 0.3 and 0.6 THz. Viability of NB1RGB and HCE-T cells after 70 h or 3 h of exposure following the MTT assay. Viability of NB1RGB and HCE-T cells subjected to different exposure conditions sham cells incubated at 43 °C for 70 h and cells cultured with 0.7 M DMSO for 3 h as a positive control. The data display the mean of three independent experiments ± standard deviation. * <span class="html-italic">p</span> &lt; 0.01 for t-test comparing the control and heating effects or with DMSO.</p>
Full article ">Figure 7
<p>Change in cell shape the before (left) and after (right) (×63), phase contrast (top) and fluorescence (bottom) micrographs. (<b>a</b>) control treatment. (<b>b</b>–<b>e</b>) 0.3, 0.4, 0.5, 0.6 THz irradiation respectively.</p>
Full article ">Figure 8
<p>Change in cell shape (<b>a</b>) with heat treatment (×20) and (<b>b</b>) with dimethylsulfoxide (DMSO) treatment (×63) before (left) and after (right), phase contrast (top) and fluorescence (bottom) micrographs.</p>
Full article ">Figure 9
<p>Level of HSP70 mRNA expression observed by RT-PCR for NB1RGB cell (<b>a</b>) and HCE-T cell (<b>b</b>). Lane M is a DNA molecule size marker. Lane 70 shows cells exposed for 70 h during cell proliferation, lane 3 shows cells exposed for 3 h following cell proliferation, and lane 0 shows sham cells. β-actin was used as an internal standard.</p>
Full article ">
8 pages, 1218 KiB  
Article
A Novel Hexahedron Photonic Crystal Fiber in Terahertz Propagation: Design and Analysis
by Bikash Kumar Paul, Md. Ashraful Haque, Kawsar Ahmed and Shuvo Sen
Photonics 2019, 6(1), 32; https://doi.org/10.3390/photonics6010032 - 21 Mar 2019
Cited by 46 | Viewed by 4389
Abstract
A novel hexahedron fiber has been proposed for biomedical imaging applications and efficient guiding of terahertz radiation. A finite element method (FEM) has been applied to investigate the guiding properties rigorously. All numerically computational investigated results for optimum parameters have revealed the high [...] Read more.
A novel hexahedron fiber has been proposed for biomedical imaging applications and efficient guiding of terahertz radiation. A finite element method (FEM) has been applied to investigate the guiding properties rigorously. All numerically computational investigated results for optimum parameters have revealed the high numerical aperture (NA) of 0.52, high core power fraction of 64%, near zero flattened dispersion of 0.5 ± 0.6 ps/THz/cm over the 0.8–1.4 THz band and low losses with 80% of the bulk absorption material loss. In addition, the V–parameter is also inspected for checking the proposed fiber modality. The proposed single-mode hexahedron photonic crystal fiber (PCF) can be highly applicable for convenient broadband transmission and numerous applications in THz technology. Full article
(This article belongs to the Special Issue Terahertz Photonics)
Show Figures

Figure 1

Figure 1
<p>Schematic view of proposed PCF with modes field for different porosities.</p>
Full article ">Figure 2
<p>The numerical aperture as a function of frequency and porosities.</p>
Full article ">Figure 3
<p>The effective area of optimum structure as a function of frequency.</p>
Full article ">Figure 4
<p>The confinement loss and scattering loss as a function of frequency.</p>
Full article ">Figure 5
<p>The effective material loss (EML) as a function of frequency for different porosities.</p>
Full article ">Figure 6
<p>The core power fraction as a function of frequency for different porosities.</p>
Full article ">Figure 7
<p>The V<math display="inline"><semantics> <msub> <mrow/> <mrow> <mi>e</mi> <mi>f</mi> <mi>f</mi> </mrow> </msub> </semantics></math> as a function of frequency for different porosities.</p>
Full article ">Figure 8
<p>The dispersion of different porosities as a function of frequency.</p>
Full article ">
Article
Long Wavelength (λ > 17 µm) Distributed Feedback Quantum Cascade Lasers Operating in a Continuous Wave at Room Temperature
by Hoang Nguyen Van, Zeineb Loghmari, Hadrien Philip, Michael Bahriz, Alexei N. Baranov and Roland Teissier
Photonics 2019, 6(1), 31; https://doi.org/10.3390/photonics6010031 - 21 Mar 2019
Cited by 21 | Viewed by 5538
Abstract
The extension of the available spectral range covered by quantum cascade lasers (QCL) would allow one to address new molecular spectroscopy applications, in particular in the long wavelength domain of the mid-infrared. We report in this paper the realization of distributed feedback (DFB) [...] Read more.
The extension of the available spectral range covered by quantum cascade lasers (QCL) would allow one to address new molecular spectroscopy applications, in particular in the long wavelength domain of the mid-infrared. We report in this paper the realization of distributed feedback (DFB) QCLs, made of InAs and AlSb, that demonstrated a continuous wave (CW) and a single mode emission at a wavelength of 17.7 µm, with output powers in the mW range. This is the longest wavelength for DFB QCLs, and for any QCLs or semiconductor lasers in general, operating in a CW at room temperature. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Band diagram of a portion of the studied active region. The laser transition is between the red (<span class="html-italic">up</span>) and green (<span class="html-italic">down</span>) states. Two quantum wells in the middle of the injector are doped with Si for an electron sheet density of 0.3 × 10<sup>11</sup> cm<sup>−2</sup> per stage.</p>
Full article ">Figure 2
<p>Room temperature electrical and optical characteristics of a typical FP quantum cascade laser (QCL) fabricated from the studied wafer. The laser is 3.6 mm long and 16 µm wide and driven with 330 ns current pulses at a repetition rate of 12 kHz. Inset: the pulsed emission spectrum of the laser.</p>
Full article ">Figure 3
<p>(<b>a</b>) Emission spectra in pulsed mode at 25 °C of DFB lasers with a different period of the DFB grating (Λ). The linewidth is limited by the resolution of the FTIR spectrometer. (<b>b</b>) Peak wavelength of the tested devices as a function of the period of the DFB grating.</p>
Full article ">Figure 4
<p>(<b>a</b>) Comparison of the pulsed characteristics of FP and DFB QCLs with different grating periods. The FP laser is 3.6 mm long and 16 µm wide; the DFB laser with Λ = 2.666 µm is 3.4 mm long and 15 µm wide; the laser with Λ = 2.590 µm is 3.0 mm long and 14 µm wide. (<b>b</b>) Threshold current densities of the studied devices as a function of operating temperature. The open symbols are for pulsed operation, and the solid symbols are for CW operation.</p>
Full article ">Figure 5
<p>(<b>a</b>) CW characteristics of a 3.4-mm-long and 14-µm-wide DFB QCL with Λ = 2.666 µm. The optical power is the power collected from one facet with a f/1 off-axis parabolic mirror without any correction for the collection efficiency. (<b>b</b>) Emission spectra in a CW as a function of the sample holder temperature for a current of 700 mA. The linewidth is limited by the resolution of the FTIR spectrometer.</p>
Full article ">Figure 6
<p>(<b>a</b>) CW characteristics of a 3.0-mm-long and 14-µm-wide DFB QCL with Λ = 2.590 µm, mounted in a Peltier cooled module. The optical power is the power collected from one facet with a f/1 off-axis parabolic mirror without any correction for the collection efficiency. (<b>b</b>) Emission spectra in a CW as a function of the sample holder temperature, for a set of currents starting at the maximum current of 580 mA and decreasing in 20 mA steps. The linewidth is limited by the resolution of the FTIR spectrometer.</p>
Full article ">
13 pages, 2441 KiB  
Article
Design an All-Optical Combinational Logic Circuits Based on Nano-Ring Insulator-Metal-Insulator Plasmonic Waveguides
by Saif Hasan Abdulnabi and Mohammed Nadhim Abbas
Photonics 2019, 6(1), 30; https://doi.org/10.3390/photonics6010030 - 19 Mar 2019
Cited by 40 | Viewed by 4440
Abstract
In this paper, we propose, analyze and simulate a new configuration to simulate all-optical combinational logic functions based on Nano-rings insulator-metal-insulator (IMI) plasmonic waveguides. We used Finite Element Method (FEM) to analyze the proposed plasmonic combinational logic functions. The analyzed combinational logic functions [...] Read more.
In this paper, we propose, analyze and simulate a new configuration to simulate all-optical combinational logic functions based on Nano-rings insulator-metal-insulator (IMI) plasmonic waveguides. We used Finite Element Method (FEM) to analyze the proposed plasmonic combinational logic functions. The analyzed combinational logic functions are Half-Adder, Full-Adder, Half-Subtractor, and Comparator One-Bit. The operation principle of these combinational logic functions is based on the constructive and destructive interferences between the input signal(s) and control signal. Numerical simulations show that a transmission threshold exists (0.25) which allows all proposed four plasmonic combinational logic functions to be achieved in one structure. As a result, the transmission threshold value measures the performance of the proposed plasmonic combinational logic functions. We use the same structure with the same dimensions at 1550 nm wavelength for all proposed plasmonic combinational logic functions. The proposed all-optical combinational logic functions structure contributes significantly to photonic integrated circuits construction and all-optical signal processing nano-circuits. Full article
Show Figures

Figure 1

Figure 1
<p>The proposed structure for the proposed plasmonic four combinational logic functions.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) the conventional half-adder logic diagram and its truth table, respectively. (<b>c</b>) The transmission spectrum of the proposed plasmonic half-adder for different states, according to its truth table. (<b>d</b>,<b>e</b>) the magnetic field distribution of Logic 00 and Logic 11 inputs, respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) the conventional half-subtractor logic diagram and its truth table, respectively. (<b>c</b>) The transmission spectrum of the proposed plasmonic half-subtractor for different states, according to its truth table. (<b>d</b>,<b>e</b>) the magnetic field distribution of Logic 10 and Logic 11 inputs, respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>) and (<b>b</b>) the conventional comparator one-bit logic diagram and its truth table, respectively. (<b>c</b>) The transmission spectrum of the proposed plasmonic comparator one-bit for different states, according to its truth table. (<b>d</b>), (<b>e</b>), (<b>f</b>) and (<b>g</b>) the magnetic field distribution of Logic 00, 01, 10, and 11 inputs, respectively.</p>
Full article ">Figure 5
<p>(<b>a</b>) and (<b>b</b>) the conventional full-adder logic symbol and its truth table, respectively. (<b>c</b>) The transmission spectrum of the proposed plasmonic full-adder for different states, according to its truth table. (<b>d</b>), (<b>e</b>), (<b>f</b>) and (<b>g</b>) the magnetic field distribution of Logic 001, 011, 110, and 111 inputs, respectively.</p>
Full article ">
8 pages, 6783 KiB  
Article
Tuning Plasmon Induced Reflectance with Hybrid Metasurfaces
by Mohsin Habib, Ekmel Ozbay and Humeyra Caglayan
Photonics 2019, 6(1), 29; https://doi.org/10.3390/photonics6010029 - 16 Mar 2019
Cited by 2 | Viewed by 3256
Abstract
Electrically tunable metasurfaces with graphene offer design flexibility to efficiently manipulate and control light. These metasurfaces can be used to generate plasmon-induced reflectance (PIR), which can be tuned by electrostatic doping of the graphene layer. We numerically investigated two designs for tunable PIR [...] Read more.
Electrically tunable metasurfaces with graphene offer design flexibility to efficiently manipulate and control light. These metasurfaces can be used to generate plasmon-induced reflectance (PIR), which can be tuned by electrostatic doping of the graphene layer. We numerically investigated two designs for tunable PIR devices using the finite difference time-domain (FDTD) method. The first design is based on two rectangular antennas of the same size and a disk; in the second design, two parallel rectangular antennas with different dimensions are used. The PIR-effect was achieved by weak hybridization of two bright modes in both devices and tuned by changing the Fermi level of graphene. A total shift of ∼362 nm was observed in the design with the modulation depth of 53% and a spectral contrast ratio of 76%. These tunable PIR devices can be used for tunable enhanced biosensing and switchable systems. Full article
Show Figures

Figure 1

Figure 1
<p>The unit cell of (<b>a</b>) the first device with two rectangular antennas and a disk. (<b>b</b>) The second device with two parallel rectangular antennas. (<b>c</b>,<b>d</b>) 3D schematics of the first and second devices, respectively.</p>
Full article ">Figure 2
<p>Simulated reflection results of (<b>a</b>) the first device and (<b>b</b>) the second device at 0.2, 0.5, and 0.8 eV . The insets show the reflection spectra of the bright modes excited individually: (<b>a</b>) shorter rectangular antenna (pink), longer antenna (green), and (<b>b</b>) rectangular antennas (pink) and disk (green).</p>
Full article ">Figure 3
<p>Electric-field (E-field) magnitudes at 0.5 eV: (<b>a</b>) E-field magnitude at 3.62 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m. (<b>b</b>) E-field magnitude at 3.94 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m. (<b>c</b>) E-field magnitude at 4.39 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m.</p>
Full article ">Figure 4
<p>E-field magnitudes at 0.5 eV: (<b>a</b>) E-field magnitude at 3.49 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m. (<b>b</b>) E-field magnitude at 3.80 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m. (<b>c</b>) E-field magnitude at 4.22 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m.</p>
Full article ">Figure 5
<p>Reflection results of the two rectangular antenna design. The SEM image is presented in the inset.</p>
Full article ">
12 pages, 4081 KiB  
Article
Epsilon-Near-Zero Absorber by Tamm Plasmon Polariton
by Rashid G. Bikbaev, Stepan Ya. Vetrov and Ivan V. Timofeev
Photonics 2019, 6(1), 28; https://doi.org/10.3390/photonics6010028 - 9 Mar 2019
Cited by 33 | Viewed by 5137
Abstract
Two schemes of excitation of a Tamm plasmon polariton localized at the interface between a photonic crystal and a nanocomposite with near-zero effective permittivity have been investigated in the framework of the temporal coupled-mode theory. The parameters of the structure have been determined, [...] Read more.
Two schemes of excitation of a Tamm plasmon polariton localized at the interface between a photonic crystal and a nanocomposite with near-zero effective permittivity have been investigated in the framework of the temporal coupled-mode theory. The parameters of the structure have been determined, which correspond to the critical coupling of the incident field with a Tamm plasmon polariton and, consequently, ensure the total absorption of the incident radiation by the structure. It has been established that the spectral width of the absorption line depends on the scheme of Tamm plasmon polariton excitation and the parameters of a nanocomposite film. The features of field localization at the Tamm plasmon polariton frequency for different excitation schemes have been examined. It has been demonstrated that such media can be used as narrowband absorbers based on Tamm plasmon polaritons localized at the interface between a photonic crystal and a nanocomposite with near-zero effective permittivity. Full article
(This article belongs to the Special Issue Advanced Optical Materials and Devices)
Show Figures

Figure 1

Figure 1
<p>Schematic of a one-dimensional photonic crystal (PhC) coupled with a nanocomposite (NC) layer with the near-zero permittivity.</p>
Full article ">Figure 2
<p>Dependences of the real <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mspace width="3.33333pt"/> <msub> <mi>ε</mi> <mi mathvariant="italic">eff</mi> </msub> </mrow> </semantics></math> (purple line) and imaginary <math display="inline"><semantics> <mrow> <mi>I</mi> <mi>m</mi> <mspace width="3.33333pt"/> <msub> <mi>ε</mi> <mi mathvariant="italic">eff</mi> </msub> </mrow> </semantics></math> (green line) parts of the effective permittivity <math display="inline"><semantics> <msub> <mi>ε</mi> <mi mathvariant="italic">eff</mi> </msub> </semantics></math> on the incident light wavelength. The filling factor is <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>11</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Schematic of the Tamm plasmon polariton (TPP) excitation from the side of (<b>a</b>) an NC film and (<b>b</b>) a PhC.</p>
Full article ">Figure 4
<p>(<b>a</b>) Transmission and (<b>b</b>) absorption spectra of the NC film at different film thicknesses and incident radiation wavelengths. The filling factor is <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>11</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>The dependence of absorptance of the NC film on its transmittance at different film thicknesses (blue line) and the dependence of the absorptance of an opaque NC film on the PhC transmittance at different numbers of PhC periods (red line). The point of intersection between the black and blue lines corresponds to the critical coupling condition (<a href="#FD12-photonics-06-00028" class="html-disp-formula">12</a>), and the point of intersection between the black and red lines corresponds to Condition (<a href="#FD13-photonics-06-00028" class="html-disp-formula">13</a>). The wavelength is <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>407</mn> <mo>.</mo> <mn>1</mn> </mrow> </semantics></math> nm.</p>
Full article ">Figure 6
<p>(<b>a</b>) NC-PhC and (<b>b</b>) PhC-NC reflectance spectra of the structure at different NC layer thicknesses <math display="inline"><semantics> <msub> <mi>d</mi> <mi mathvariant="italic">eff</mi> </msub> </semantics></math> and a constant filling factor of <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>11</mn> </mrow> </semantics></math>. The green dashed line shows the reflectance spectra of the structures under critical coupling conditions.</p>
Full article ">Figure 7
<p>Transmittance (red line), reflectance (blue line), and absorptance (black line) spectra of the structure under the critical coupling conditions upon TPP excitation from (<b>a</b>) the NC and (<b>b</b>) PhC side. The NC layer thickness and number of PhC periods are <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mi mathvariant="italic">eff</mi> </msub> <mo>=</mo> <mn>201</mn> </mrow> </semantics></math> nm, <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>25</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mi mathvariant="italic">eff</mi> </msub> <mo>=</mo> <mn>700</mn> </mrow> </semantics></math> nm, <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>, respectively.</p>
Full article ">Figure 8
<p>TPP in the region of positive <math display="inline"><semantics> <msub> <mi>ε</mi> <mi mathvariant="italic">eff</mi> </msub> </semantics></math>. Spatial distribution of the refractive index of the structure (green line) and local field intensity at the TPP wavelength (red line) for the TPP excitation from (<b>a</b>) the NC and (<b>b</b>) PhC side.</p>
Full article ">Figure 9
<p>(<b>a</b>) TPP in the metal-like NC region <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mi mathvariant="italic">eff</mi> </msub> <mo>&lt;</mo> <mn>0</mn> </mrow> </semantics></math>. Transmittance (black line), reflectance (blue line), and absorptance (red line) spectra of the structure under the critical coupling conditions for the scheme of TPP excitation from the PhC side at a thickness of <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mi mathvariant="italic">first</mi> </msub> <mo>=</mo> <mn>78</mn> </mrow> </semantics></math> nm of the layer adjacent to the PhC and (<b>b</b>) the spatial distribution of the local field intensity at the TPP wavelength.</p>
Full article ">
12 pages, 2397 KiB  
Article
Mode Suppression in Injection Locked Multi-Mode and Single-Mode Lasers for Optical Demultiplexing
by Kevin Shortiss, Maryam Shayesteh, William Cotter, Alison H. Perrott, Mohamad Dernaika and Frank H. Peters
Photonics 2019, 6(1), 27; https://doi.org/10.3390/photonics6010027 - 8 Mar 2019
Cited by 13 | Viewed by 3950
Abstract
Optical injection locking has been demonstrated as an effective filter for optical communications. These optical filters have advantages over conventional passive filters, as they can be used on active material, allowing them to be monolithically integrated onto an optical circuit. We present an [...] Read more.
Optical injection locking has been demonstrated as an effective filter for optical communications. These optical filters have advantages over conventional passive filters, as they can be used on active material, allowing them to be monolithically integrated onto an optical circuit. We present an experimental and theoretical study of the optical suppression in injection locked Fabry–Pérot and slotted Fabry–Pérot lasers. We consider both single frequency and optical comb injection. Our model is then used to demonstrate that improving the Q factor of devices increases the suppression obtained when injecting optical combs. We show that increasing the Q factor while fixing the device pump rate relative to threshold causes the locking range of these demultiplexers to asymptotically approach a constant value. Full article
(This article belongs to the Special Issue Semiconductor Laser Dynamics: Fundamentals and Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Illustration of a photonic integrated circuit for demultiplexing optical combs. The comb is first split equally using a multimode interferometer, and then, individual slave lasers are frequency locked to specific lines in the comb. (<b>b</b>) Illustrations of the fields inside a laser cavity, with reflecting mirrors <math display="inline"><semantics> <msub> <mi>r</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>r</mi> <mn>2</mn> </msub> </semantics></math>.</p>
Full article ">Figure 2
<p>Comparison between measured device gain (left vertical axis) from a Fabry–Pérot (FP) laser calculated using the Cassidy gain method [<a href="#B29-photonics-06-00027" class="html-bibr">29</a>] and the gain <math display="inline"><semantics> <msub> <mi>g</mi> <mi>m</mi> </msub> </semantics></math> implemented in the model (right vertical axis).</p>
Full article ">Figure 3
<p>Setup used to measure the intensity plots of the optical injection locking experiments. Dashed lines indicate the additional setup used when injecting optical combs. TLS: tunable laser source, MZM: Mach–Zehnder modulator, RF Gen: RF Generator, Iso: Isolator, PC: polarisation controller, EDFA: erbium–doped fibre amplifier, OSA: optical spectrum analyser, DUT: device under test.</p>
Full article ">Figure 4
<p>Experimental and calculated injected wavelength sweeps of a 700-<math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m FP device. In both cases, the slave laser was biased at 2.5-times the threshold. (<b>a</b>) Experimental sweep, for an injected power of <math display="inline"><semantics> <mrow> <mo>-</mo> <mn>12</mn> <mo>.</mo> <mn>5</mn> </mrow> </semantics></math> dBm and free running slave power of <math display="inline"><semantics> <mrow> <mo>-</mo> <mn>4</mn> </mrow> </semantics></math> dBm. (<b>b</b>) Calculated sweep, for an injection ratio of <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>.</mo> <mn>33</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>-</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Experimental and calculated temperature sweeps of an optically-injected 600 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m-long two-section slotted FP device, with a single etched slot in the centre of the device separating the sections. In each case, the slave laser was under optical injection at a wavelength of 1563.36 nm. (<b>a</b>) Experimental sweep from [<a href="#B12-photonics-06-00027" class="html-bibr">12</a>]. (<b>b</b>) Calculated sweep for an injection ratio of <math display="inline"><semantics> <mrow> <mn>6</mn> <mo>.</mo> <mn>13</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>-</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>(<b>a</b>) Comb demultiplexer, featuring a <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>×</mo> <mn>2</mn> </mrow> </semantics></math> multimode interferometer (MMI) and two SFP lasers [<a href="#B25-photonics-06-00027" class="html-bibr">25</a>]. (<b>b</b>) Optical comb injected into the demultiplexer. This two line comb was generated by biasing the MZM at the point where the carrier is suppressed, giving two strong lines.</p>
Full article ">Figure 7
<p>(<b>a</b>) Experimental and (<b>b</b>) calculated injected comb sweeps of an optical demultiplexer, as shown in <a href="#photonics-06-00027-f006" class="html-fig">Figure 6</a>a, with an injected optical comb as shown in <a href="#photonics-06-00027-f006" class="html-fig">Figure 6</a>b. Due to the optical coupling of the lensed fibre, both outputs of the multiplexer could not be measured simultaneously.</p>
Full article ">Figure 8
<p>Calculated results from optical comb injection simulations, of a 700-<math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m FP laser. (<b>a</b>) Plot showing how the SMSR of the output spectrum varies as the Q of the laser cavity increases, for two different injection ratios, assuming zero detuning and biased at 3.0-times the threshold. The higher injection ratio was initialised at <math display="inline"><semantics> <mrow> <mn>16</mn> <mo>.</mo> <mn>9</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>-</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math> and the corresponding lower injection ratio at <math display="inline"><semantics> <mrow> <mn>10</mn> <mo>.</mo> <mn>1</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>-</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math>. For qualitative comparison, the side mode suppression ratio (SMSR) from a passive cavity with equivalent Q is also plotted. (<b>b</b>) Intensity plot of how SMSR varies versus detuning and the Q factor, for an injection ratio of <math display="inline"><semantics> <mrow> <mn>6</mn> <mo>.</mo> <mn>7</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>-</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math>, at a current of 2.5-times the threshold. The white regions indicate where the slave laser was unlocked. (<b>c</b>) Plot of the locking range of the FP laser versus Q, for the same injection ratios and parameters as in (<b>a</b>).</p>
Full article ">
11 pages, 2574 KiB  
Article
The Design of Optical Circuit-Analog Absorbers through Electrically Small Nanoparticles
by Alessio Monti, Andrea Alù, Alessandro Toscano and Filiberto Bilotti
Photonics 2019, 6(1), 26; https://doi.org/10.3390/photonics6010026 - 6 Mar 2019
Cited by 10 | Viewed by 4147
Abstract
In the last few years, the perfect absorption of light has become an important research topic due to its dramatic impact in photovoltaics, photodetectors, color filters and thermal emitters. While broadband optical absorption is relatively easy to achieve using bulky devices, today there [...] Read more.
In the last few years, the perfect absorption of light has become an important research topic due to its dramatic impact in photovoltaics, photodetectors, color filters and thermal emitters. While broadband optical absorption is relatively easy to achieve using bulky devices, today there is a strong need and interest in achieving the same effects by employing nanometric structures that are compatible with modern nanophotonic components. In this paper, we propose a general procedure to design broadband nanometer-scale absorbers working in the optical spectrum. The proposed devices, which can be considered an extension to optics of microwave circuit-analog absorbers, consist of several layers containing arrays of elongated nanoparticles, whose dimensions are engineered to control both the absorption level and the operational bandwidth. By combining a surface-impedance homogenization and an equivalent transmission-line formalism, we define a general analytical procedure that can be employed to achieve a final working design. As a relevant example, we show that the proposed approach allows designing an optical absorber exhibiting a 20% fractional bandwidth on a thickness of λ/4 at the central frequency of operation. Full-wave results confirming the effectiveness of the analytical findings, as well as some considerations about the experimental realization of the proposed devices are provided. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) A sketch of a three-layered circuit-analog absorber and (<b>b</b>) its equivalent circuit model.</p>
Full article ">Figure 2
<p>(<b>a</b>) The optical metasurface made by an array of ellipsoidal plasmonic nanoparticles arranged in a square lattice; (<b>b</b>) the complex surface impedance of two different optical metasurfaces designed to have the same surface resistance (<span class="html-italic">R<sub>s</sub></span> = 200 Ω) and opposite reactive behaviors (capacitive and inductive). The ticks represent the results of full-wave simulations.</p>
Full article ">Figure 3
<p>(<b>a</b>) The two-layered optical circuit-analogue (CA) absorber. The ground plane is replaced by a thick enough layer of silver; (<b>b</b>) the surface reactance of the two metasurfaces required to maximize the absorption within the frequencies range 550–650 THz. The surface reactances of the metasurfaces are compared with the input reactance of the absorber evaluated before the first and the second layer.</p>
Full article ">Figure 4
<p>(<b>a</b>) The input impedance of the designed CA absorber and of an equivalent Salisbury screen with the same thickness. (<b>b</b>) The analytical and numerical absorbance (absolute value) of the designed CA absorber and of an equivalent Salisbury screen with the same thickness. Ticks represent the results of full-wave simulations.</p>
Full article ">Figure 5
<p>(<b>a</b>) The absorbance, reflectance and transmittance (absolute value) of a three-layered CA optical absorber obtained through full-wave simulations; (<b>b</b>) the value of the absorbance (absolute value) at three different frequencies as the impinging angle changes.</p>
Full article ">Figure 6
<p>(<b>a</b>) The complex surface impedance exhibited by two arrays of nanoellipsoids and nanocylinders with the same size; (<b>b</b>) the absorbance of the three-layered CA absorber made by nanocylinders for different values of the perturbation parameter Δ. In the inset, the equivalent absorber structure for Δ = 6 nm is shown.</p>
Full article ">
7 pages, 1971 KiB  
Article
Numerical Study on the Soliton Mode-Locking of the Er3+-Doped Fluoride Fiber Laser at ~3 μm with Nonlinear Polarization Rotation
by Feijuan Zhang, Wenyan Yan, Shengnan Liang, Chao Tan and Pinghua Tang
Photonics 2019, 6(1), 25; https://doi.org/10.3390/photonics6010025 - 6 Mar 2019
Cited by 5 | Viewed by 3806
Abstract
Recent interest in the application of mid-infrared (mid-IR) lasers has made the generation of ~3 µm ultrafast pulses a hot topic. Recently, the generation of femtosecond-scale pulses in Er3+-doped fluoride fiber lasers has been realized by nonlinear polarization rotation (NPR). However, [...] Read more.
Recent interest in the application of mid-infrared (mid-IR) lasers has made the generation of ~3 µm ultrafast pulses a hot topic. Recently, the generation of femtosecond-scale pulses in Er3+-doped fluoride fiber lasers has been realized by nonlinear polarization rotation (NPR). However, a numerical study on these fiber lasers has not been reported yet. In this work, the output properties of the NPR passively mode-locked Er3+-doped fluoride fiber ring laser in ~3 µm have been numerically investigated based on the coupled Ginzburg–Landu equation. The simulation results indicate that stable uniform solitons (0.75 nJ) with the pulse duration of femtosecond-scale can be generated from this fiber laser. This numerical investigation can provide some reference for developing the high energy femtosecond soliton fiber lasers in the mid-IR. Full article
(This article belongs to the Special Issue Fiber Lasers)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the passively mode-locked Er<sup>3+</sup>-doped fluoride fiber laser. PD-ISO, polarization-dependent isolator; QWP, quarter waveplate; HWP, half waveplate; DM, dichroic mirror; BS, beam splitter.</p>
Full article ">Figure 2
<p>Cavity transmissions with respect to the total phase delay.</p>
Full article ">Figure 3
<p>Pulse (<b>a</b>) and spectrum (<b>b</b>) evolution with round trips. Output pulse profile (<b>c</b>) and spectrum profile (<b>d</b>) in the 120th round trip.</p>
Full article ">Figure 4
<p>Output pulse characteristics of the passively mode-locked Er<sup>3+</sup>-doped fluoride fiber laser. (<b>a</b>) Pulse width and 3 dB spectral bandwidth and (<b>b</b>) peak power and pulse energy as a function of fiber length.</p>
Full article ">Figure 5
<p>Output pulse characteristics of the passively mode-locked Er<sup>3+</sup>-doped fluoride fiber laser. (<b>a</b>) Pulse width and 3 dB spectral bandwidth and (<b>b</b>) peak power and pulse energy as a function of small signal gain.</p>
Full article ">Figure 6
<p>Pulse evolution for <span class="html-italic">g<sub>0</sub></span> = 1.15 m<sup>−1</sup> (<b>a</b>) and 1.5 m<sup>−1</sup> (<b>b</b>) with round trips.</p>
Full article ">
18 pages, 1926 KiB  
Review
Recent Progress on Ge/SiGe Quantum Well Optical Modulators, Detectors, and Emitters for Optical Interconnects
by Papichaya Chaisakul, Vladyslav Vakarin, Jacopo Frigerio, Daniel Chrastina, Giovanni Isella, Laurent Vivien and Delphine Marris-Morini
Photonics 2019, 6(1), 24; https://doi.org/10.3390/photonics6010024 - 1 Mar 2019
Cited by 29 | Viewed by 7570
Abstract
Germanium/Silicon-Germanium (Ge/SiGe) multiple quantum wells receive great attention for the realization of Si-based optical modulators, photodetectors, and light emitters for short distance optical interconnects on Si chips. Ge quantum wells incorporated between SiGe barriers, allowing a strong electro-absorption mechanism of the quantum-confined Stark [...] Read more.
Germanium/Silicon-Germanium (Ge/SiGe) multiple quantum wells receive great attention for the realization of Si-based optical modulators, photodetectors, and light emitters for short distance optical interconnects on Si chips. Ge quantum wells incorporated between SiGe barriers, allowing a strong electro-absorption mechanism of the quantum-confined Stark effect (QCSE) within telecommunication wavelengths. In this review, we respectively discuss the current state of knowledge and progress of developing optical modulators, photodetectors, and emitters based on Ge/SiGe quantum wells. Key performance parameters, including extinction ratio, optical loss, swing bias voltages, and electric fields, and modulation bandwidth for optical modulators, dark currents, and optical responsivities for photodetectors, and emission characteristics of the structures will be presented. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic view of side-entry Ge/SiGe MQW optical modulator [<a href="#B36-photonics-06-00024" class="html-bibr">36</a>]. (<b>b</b>,<b>c</b>) Schematic views of vertical-incidence electro-absorption modulators [<a href="#B39-photonics-06-00024" class="html-bibr">39</a>,<a href="#B40-photonics-06-00024" class="html-bibr">40</a>]. (Reproduced with permission from [<a href="#B36-photonics-06-00024" class="html-bibr">36</a>] and [<a href="#B39-photonics-06-00024" class="html-bibr">39</a>] © 2007 and 2012 the Optical Society, and [<a href="#B40-photonics-06-00024" class="html-bibr">40</a>] © 2013 IEEE).</p>
Full article ">Figure 2
<p>(<b>a</b>) QCSE from a Ge/SiGe MQW 34-µm-long planar waveguide obtained from optical transmission measurements [<a href="#B43-photonics-06-00024" class="html-bibr">43</a>], and (<b>b</b>) a stand-alone Ge/SiGe MQW waveguide optical modulator. (<b>c</b>) Scanning electron microscope (SEM) image of the stand-alone optical modulator [<a href="#B46-photonics-06-00024" class="html-bibr">46</a>]. (<b>d</b>) Schematic view of SOI-waveguide-integrated Ge/SiGe MQW optical modulator using a butt coupling approach [<a href="#B65-photonics-06-00024" class="html-bibr">65</a>]. (<b>e</b>) Schematic and SEM views of a SiGe waveguide-integrated Ge/SiGe MQW optical modulator using a linear taper coupling approach [<a href="#B68-photonics-06-00024" class="html-bibr">68</a>] (Reproduced with permission from [<a href="#B43-photonics-06-00024" class="html-bibr">43</a>] and [<a href="#B46-photonics-06-00024" class="html-bibr">46</a>] © 2011 and 2012 the Optical Society, [<a href="#B65-photonics-06-00024" class="html-bibr">65</a>] © 2011 IEEE, and [<a href="#B68-photonics-06-00024" class="html-bibr">68</a>] © 2014 Springer Nature).</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic view of a cross section of Ge/SiGe MQW waveguide photodetector, and in the inset, an optical microscope top view of the waveguide photodetector. (<b>b</b>) The 10 Gb/s electrical outputs of a 80-µm-long Ge/SiGe MQW photodetector. (<b>c</b>) Optical responsivity of the Ge/SiGe MQW photodetector spectra at different reverse bias (reprinted with permission from [<a href="#B75-photonics-06-00024" class="html-bibr">75</a>] © 2011 IEEE). (<b>d</b>) Schematic and (<b>e</b>) SEM views of a surface-illuminated Ge/SiGe MQW photodiode (12-µm diameter).</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic view of scattering, thermalization, and recombination processes that occur in Ge/SiGe QWs of excited electrons [<a href="#B94-photonics-06-00024" class="html-bibr">94</a>]. (<b>b</b>) Photoluminescence spectra (full lines) and absorption spectra (dashed lines) of Ge/SiGe MQWs measured at room temperature and at 5 K [<a href="#B95-photonics-06-00024" class="html-bibr">95</a>]. (<b>c</b>) EL spectra from 80-µm-long waveguide light-emitting-diode, with and without an optical polarizer between the waveguide output and the detector, and SEM view of Ge/SiGe MQW light emitting diode in a waveguide structure. EL spectra were observed to be TE polarized [<a href="#B98-photonics-06-00024" class="html-bibr">98</a>]. (<b>d</b>) EL spectra at different temperatures under a constant injection current [<a href="#B109-photonics-06-00024" class="html-bibr">109</a>]. (Reproduced with permission from [<a href="#B94-photonics-06-00024" class="html-bibr">94</a>] © 2011 American Physical Society, [<a href="#B95-photonics-06-00024" class="html-bibr">95</a>,<a href="#B98-photonics-06-00024" class="html-bibr">98</a>] © 2011 AIP Publishing, and [<a href="#B109-photonics-06-00024" class="html-bibr">109</a>] © MDPI under Creative Commons Attribution (CC-BY) license).</p>
Full article ">
11 pages, 2077 KiB  
Article
Photonic Inverse Design of Simple Particles with Realistic Losses in the Visible Frequency Range
by Constantinos Valagiannopoulos
Photonics 2019, 6(1), 23; https://doi.org/10.3390/photonics6010023 - 28 Feb 2019
Cited by 5 | Viewed by 4057
Abstract
Billions of U.S. dollars of basic and applied research funding have been invested during the last few years in ideas proposing inverse concepts. The photonics market could not make an exception to this global trend, and thus, several agenda-setting research groups have already [...] Read more.
Billions of U.S. dollars of basic and applied research funding have been invested during the last few years in ideas proposing inverse concepts. The photonics market could not make an exception to this global trend, and thus, several agenda-setting research groups have already started providing sophisticated tools, constrained optimization algorithms, and selective evolution techniques towards this direction. Here, we present an approach of inverse design based on the exhaustive trial-and-testing of the available media and changing the physical dimensions’ range according to the operational wavelength. The proposed technique is applied to the case of an optimal radiation-enhancing cylindrical particle fed by a line source of visible light and gives a two-order increase in the magnitude of the produced signal. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic diagram of the forward and inverse design of an electromagnetic system (reproduced with permission from [<a href="#B7-photonics-06-00023" class="html-bibr">7</a>]); (<b>b</b>) illustrative description of the optimization process sweeping the model parameters.</p>
Full article ">Figure 2
<p>(<b>a</b>) Approximate loci of the complex permittivities of basic metals and chemical compounds for various frequencies of visible light (reproduced with permission from [<a href="#B7-photonics-06-00023" class="html-bibr">7</a>]); (<b>b</b>) exact dispersion of the real part of permittivities across the visible spectrum of the used materials; (<b>c</b>) physical configuration of the considered radiation-enhancing particles.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) Maximum relative radiation as function of operational wavelength <math display="inline"><semantics> <mi>λ</mi> </semantics></math> with use of covers made of several materials for: (<b>a</b>) TM excitation and (<b>b</b>) TE excitation; (<b>c</b>,<b>d</b>) trajectories of optimal designs on the <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>a</mi> <mo>/</mo> <mi>b</mi> <mo>,</mo> <mi>b</mi> <mo>/</mo> <mi>λ</mi> <mo>)</mo> </mrow> </semantics></math> plane as the wavelength <math display="inline"><semantics> <mi>λ</mi> </semantics></math> changes for: (<b>c</b>) TM excitation of the systems of (<b>a</b>) and (<b>c</b>) TE excitation of the systems of (<b>b</b>).</p>
Full article ">Figure 4
<p>Spatial distribution of the normalized electric field <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mi>z</mi> </msub> <mrow> <mo>(</mo> <mi>z</mi> <mo>,</mo> <mi>y</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for: free space (left), optimal Ag cladding (center), and arbitrary, but close to optimal Ag cladding (right). The optimal design corresponds to the point of <a href="#photonics-06-00023-f003" class="html-fig">Figure 3</a>a’s curve at <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>693</mn> <mspace width="3.33333pt"/> <mi>nm</mi> </mrow> </semantics></math>. Design parameters: <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>/</mo> <mi>b</mi> <mo>=</mo> <mn>0.82</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>/</mo> <mi>λ</mi> <mo>=</mo> <mn>0.42</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Spatial distribution of the normalized magnetic field <math display="inline"><semantics> <mrow> <msub> <mi>H</mi> <mi>z</mi> </msub> <mrow> <mo>(</mo> <mi>z</mi> <mo>,</mo> <mi>y</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for: free space (<b>left</b>), optimal GaP cladding (<b>center</b>), and arbitrary, but close to optimal GaP cladding (<b>right</b>). The optimal design corresponds to the point of <a href="#photonics-06-00023-f003" class="html-fig">Figure 3</a>b’s curve at <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>460</mn> <mspace width="3.33333pt"/> <mi>nm</mi> </mrow> </semantics></math>. Design parameters: <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>/</mo> <mi>b</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>/</mo> <mi>λ</mi> <mo>=</mo> <mn>0.10</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Linear plots of the normalized signal (field squared in arbitrary units) as a function of electrical radial distance <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>/</mo> <mi>λ</mi> </mrow> </semantics></math> for: (<b>a</b>) the optimal dimensions of the TM Ag-based design (of <a href="#photonics-06-00023-f004" class="html-fig">Figure 4</a>) when the shell is empty (vacuum), silver (optimal), and a material with <math display="inline"><semantics> <mi>ε</mi> </semantics></math> of the opposite real part than that of silver; (<b>b</b>) the optimal dimensions of the TE GaP-based design (of <a href="#photonics-06-00023-f005" class="html-fig">Figure 5</a>) when the shell is empty (vacuum), GaP (optimal), and material with <math display="inline"><semantics> <mi>ε</mi> </semantics></math> of the opposite real part than that of gallium phosphide. Vertical dashed lines denote the boundaries of the radiation-enhancing shell.</p>
Full article ">Figure 7
<p>Relative radiative power <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mrow> <mi>r</mi> <mi>a</mi> <mi>d</mi> </mrow> </msub> <mo>/</mo> <msub> <mi>P</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> variation with respect to changes in aspect ratio <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>/</mo> <mi>b</mi> </mrow> </semantics></math> and electrical size <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>/</mo> <mi>λ</mi> </mrow> </semantics></math> around the optimal designs (marked by ×) of: (<b>a</b>) TM waves of the Ag cladding of <a href="#photonics-06-00023-f004" class="html-fig">Figure 4</a> at <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>693</mn> <mspace width="3.33333pt"/> <mi>nm</mi> </mrow> </semantics></math> and (<b>b</b>) TE waves of the GaP cladding of <a href="#photonics-06-00023-f005" class="html-fig">Figure 5</a> at <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>460</mn> <mspace width="3.33333pt"/> <mi>nm</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>(<b>a</b>,<b>c</b>) Radiation efficiency <math display="inline"><semantics> <msub> <mi>e</mi> <mi>R</mi> </msub> </semantics></math> in dB and (<b>b</b>,<b>d</b>) relative radiation power <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mrow> <mi>r</mi> <mi>a</mi> <mi>d</mi> </mrow> </msub> <mo>/</mo> <msub> <mi>P</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> in dB as a function of the aspect ratio <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>/</mo> <mi>b</mi> </mrow> </semantics></math> and electrical size <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>/</mo> <mi>λ</mi> </mrow> </semantics></math> for: (<b>a</b>,<b>b</b>) TM waves of Ag cladding and (<b>c</b>,<b>d</b>) TE waves of GaP cladding.</p>
Full article ">Figure 9
<p>Frequency response for both excitations (TE and TM waves) of: (<b>a</b>) the Ag cladding of <a href="#photonics-06-00023-f004" class="html-fig">Figure 4</a>, TM-optimized at <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>693</mn> <mspace width="3.33333pt"/> <mi>nm</mi> </mrow> </semantics></math>, and (<b>b</b>) the GaP cladding of <a href="#photonics-06-00023-f005" class="html-fig">Figure 5</a>, TE-optimized at <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>460</mn> <mspace width="3.33333pt"/> <mi>nm</mi> </mrow> </semantics></math>.</p>
Full article ">
27 pages, 5391 KiB  
Review
Terahertz Field Confinement in Nonlinear Metamaterials and Near-Field Imaging
by George R. Keiser and Pernille Klarskov
Photonics 2019, 6(1), 22; https://doi.org/10.3390/photonics6010022 - 28 Feb 2019
Cited by 23 | Viewed by 8805
Abstract
This article reviews recent advances in terahertz science and technology that rely on confining the energy of incident terahertz radiation to small, very sub-wavelength sized regions. We focus on two broad areas of application for such field confinement: metamaterial-based nonlinear terahertz devices and [...] Read more.
This article reviews recent advances in terahertz science and technology that rely on confining the energy of incident terahertz radiation to small, very sub-wavelength sized regions. We focus on two broad areas of application for such field confinement: metamaterial-based nonlinear terahertz devices and terahertz near-field microscopy and spectroscopy techniques. In particular, we focus on field confinement in: terahertz nonlinear absorbers, metamaterial enhanced nonlinear terahertz spectroscopy, and in sub-wavelength terahertz imaging systems. Full article
(This article belongs to the Special Issue Terahertz Photonics)
Show Figures

Figure 1

Figure 1
<p>Numerical simulations of the normalized peak (<b>a</b>) electric field and (<b>b</b>) magnetic field distributions in a split ring resonator excited by an incident THz signal. Electric fields are confined to the capacitive gap region, while magnetic fields are confined near the areas of high current density in the resonator. (<b>c</b>,<b>d</b>) Before and after pictures showing mass transfer of gold patterning in a metamaterial antenna structure. Mass transfer is induced by the highly confined resonant fields in the 1 µm thick gap region. Reproduced from [<a href="#B29-photonics-06-00022" class="html-bibr">29</a>].</p>
Full article ">Figure 2
<p>Example of structural tuning of field confinement in a split ring resonator. (<b>a</b>) Photograph of a metamaterial consisting of a SRR array below a closed conducting ring array. (<b>b</b>) Inducing an in-plane shift between the two layers, results in a 40% increase to the magnitude of the electric fields confined in the SRR gap. Reproduced from [<a href="#B36-photonics-06-00022" class="html-bibr">36</a>].</p>
Full article ">Figure 3
<p>Field confinement in nonlinear metamaterials. (<b>a</b>) A nonlinear metamaterial for second harmonic generation at microwave frequencies. A varactor in the SRR capacitive gap (lower right inset) gives rise to the nonlinear response. (<b>b</b>) THz induced optical photoluminescence due to field confinement in SRRs (low left inset). Reprinted with permission from [<a href="#B48-photonics-06-00022" class="html-bibr">48</a>,<a href="#B49-photonics-06-00022" class="html-bibr">49</a>]. Copyright 2011 and 2014, respectively, by the American Physical Society. (<b>c</b>) Actively tunable MM devices on GaAs (center inset) substrates show strong nonlinear responses due to high incident fields and field confinement. For high incident fields, the overall modulation range drops to 0 dB, due to nonlinear effects [<a href="#B50-photonics-06-00022" class="html-bibr">50</a>]. (<b>d</b>) Nonlinear MMs in the near IR can also be implemented using field confinement. Here, field confinement induces second harmonic generation of an 800 nm beam and beam shaping of the 400 nm output. Reprinted with permission from [<a href="#B55-photonics-06-00022" class="html-bibr">55</a>]. Copyright 2016 by the American Chemical Society.</p>
Full article ">Figure 4
<p>Applications of high fields and field confinement in MM detectors and saturable absorbers. (<b>a</b>) A metamaterial/pyroelectric based IR detector with enhanced sensitivity due to field confinement in the resonator gaps, where a pyroelectric material responds to the local field [<a href="#B61-photonics-06-00022" class="html-bibr">61</a>]. (<b>b</b>) Plot of the absorption and response curve for the MM from (<b>a</b>). (<b>c</b>) A THz saturable absorber designed using superconducting MMs (see inset for schematic). The resonant fields in the MM couple the nonlinear response of the YBCO superconducting SRR to produce an absorption peak that depends on the incident THz power, as shown here. Reproduced from [<a href="#B62-photonics-06-00022" class="html-bibr">62</a>]. (<b>d</b>) Another example of a THz saturable absorber. Here, small GaAs patches are placed in the confined fields of a metamaterial resonator gap to produce a power dependent absorption peak. The device is fabricated on a flexible polyimide film (see inset) [<a href="#B63-photonics-06-00022" class="html-bibr">63</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) An example of THz fields confined using a MM structure used to excite a phase and structural transition in VO<sub>2</sub>. The THz induced structural transition is probed using a 7 keV X-Ray source. Reprinted with permission from [<a href="#B84-photonics-06-00022" class="html-bibr">84</a>]. Copyright 2018 by the American Physics Society. (<b>b</b>) The induced structural transition from M1 to R phases probed in (<b>a</b>). (<b>c</b>,<b>d</b>) Confined THz fields used to excite a quantum Stark effect and photoluminescence in Cd-Se quantum dots. Reprinted with permission from [<a href="#B85-photonics-06-00022" class="html-bibr">85</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) Resonant slot antenna achieving an increased absorption cross section when reducing the slot width on the nanoscale. Reprinted with permission from [<a href="#B93-photonics-06-00022" class="html-bibr">93</a>]. (<b>b</b>) Optical pump-THz probe experiment on slot antennas showing reduced carrier life time due to the surface sensitivity of this method. Reprinted with permission from [<a href="#B96-photonics-06-00022" class="html-bibr">96</a>].</p>
Full article ">Figure 7
<p>Early work of sub-wavelength THz imaging: (<b>a</b>) Imaging setup using a tapered metal tip and (<b>b</b>) the measured test sample showing 50 µm resolution at 220 µm. Reprinted with permission from [<a href="#B101-photonics-06-00022" class="html-bibr">101</a>]. (<b>c</b>) THz SNOM setup for imaging of a silicon-gold grating and (<b>d</b>) the resulting image with a spatial resolution of 150 nm. Reprinted with permission from [<a href="#B123-photonics-06-00022" class="html-bibr">123</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) s-SNOM setup using a THz CW laser at 2.54 THz; (<b>b</b>) electro-dynamical simulations of the FC for a standard AFM probe; and (<b>c</b>) the resulting AFM and THz image. Reprinted with permission from [<a href="#B128-photonics-06-00022" class="html-bibr">128</a>].</p>
Full article ">Figure 9
<p>s-SNOM images of gold islands measured with a free-electron laser at 1.3 THz: (<b>a</b>) AFM topography; (<b>b</b>) near-field image referenced to the 2nd harmonic; and (<b>c</b>) the 3rd harmonic. (<b>d</b>) AFM topography and (<b>e</b>) the 3rd harmonic near-field image of the zoomed region indicated in (<b>d</b>). (<b>f</b>) Line scan over the gold structure indicated in (<b>e</b>). Reprinted with permission from [<a href="#B132-photonics-06-00022" class="html-bibr">132</a>]. (<b>g</b>) Illustration of THz s-SNOM imaging of the embedded gold grating and (<b>h</b>) the resulting images of AFM topography, THz peak field image and corresponding waveforms along the white line (bottom to top). Reprinted with permission from [<a href="#B134-photonics-06-00022" class="html-bibr">134</a>].</p>
Full article ">Figure 10
<p>THz photocurrent nanoscopy: (<b>a</b>) Illustration of the experimental setup. LG1 and LG2 represent the split gates of gold used to control the carrier concentration in the graphene. (<b>b</b>) Photocurrent image recorded for 2.52 THz and with n<sub>1</sub> = 0.77 × 1012 cm<sup>−2</sup> and n<sub>2</sub> = −0.71 × 1012 cm<sup>−2</sup> set by LG1 and LG2, respectively. (<b>c</b>) Near-field photocurrent profiles for different frequencies recorded along the dashed line in (<b>b</b>,<b>d</b>) Experimental (red dots) and theoretical dispersion relations in the heterostructure (blue color plot), free-standing graphene (blue solid line), light line in free space (blue dashed line), and acoustic plasmons (dashed black curve). Reprinted with permission from [<a href="#B142-photonics-06-00022" class="html-bibr">142</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) Illustration of nanoscale LTEM (Laser THz Emission Microscopy); (<b>b</b>) AFM and nanoscale LTEM image referenced at the 2nd harmonic; and (<b>c</b>) polarization dependent signals of nanoscale LTEM signal (blue), the NIR near-field signal (red) and the LTEM signal referenced to an optical chopper when turning the polarization of the incident NIR beam as illustrated in the inset. Reprinted with permission from [<a href="#B156-photonics-06-00022" class="html-bibr">156</a>].</p>
Full article ">Figure 12
<p>Resonant THz probes: (<b>a</b>) Measured THz near-field induced photocurrent as a function of antenna length (blue dots) together with full-wave simulation of the illustrated structure. The resulting simulations of the two resonances marked as the dotted black lines are shown below the plot. (<b>b</b>) Illustration and resulting field enhancements 10 nm below the apex as a function of length for four configurations: (A) free-standing tip; (B) tip with a cantilever; (C) tip with cantilever and graphene-based device as described in the text; and (D) tip with cantilever and sample device where graphene has been replaced by PEC. Reprinted with permission from [<a href="#B160-photonics-06-00022" class="html-bibr">160</a>].</p>
Full article ">Figure 13
<p>The development of sub-wavelength imaging methods over time.</p>
Full article ">
11 pages, 3455 KiB  
Article
Wedge Surface Plasmon Polariton Waveguides Based on Wet-Bulk Micromachining
by Nguyen Thanh Huong, Nguyen Van Chinh and Chu Manh Hoang
Photonics 2019, 6(1), 21; https://doi.org/10.3390/photonics6010021 - 27 Feb 2019
Cited by 12 | Viewed by 3963
Abstract
In this paper, we propose and investigate the modal characteristics of wedge surface plasmon polariton (SPP) waveguides for guiding surface plasmon waves. The wedge SPP waveguides are composed of a silver layer deposited onto the surface of a wedge-shaped silicon dielectric waveguide. The [...] Read more.
In this paper, we propose and investigate the modal characteristics of wedge surface plasmon polariton (SPP) waveguides for guiding surface plasmon waves. The wedge SPP waveguides are composed of a silver layer deposited onto the surface of a wedge-shaped silicon dielectric waveguide. The wedge-shaped silicon dielectric waveguides are explored from the anisotropic wet etching property of single crystal silicon. The wedge SPP waveguides are embedded in a dielectric medium to form the metal–dielectric interface for guiding the surface plasmon waves. The propagation characteristics of the wedge SPP waveguides at the optical telecommunication wavelength of 1.55 μm are evaluated by a numerical simulation. The influence of the physical parameters such as the dimensions of the wedge SPP waveguide and the refractive index of the dielectric medium on the propagation of the surface plasmon wave is investigated. In addition, by comparing the propagation characteristics, we derive the wedge SPP waveguide with the optimal performance. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Sketch of (<b>a</b>) the proposed wedge SPP waveguides with the sidewall corners α = 45°, 54.74°, and 90°, and(<b>b</b>) the cross-section geometry of the wedge SPP waveguide with the trapezoidal cross-section.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>d</b>) Schematic of the fabrication flow of the wedge SPP waveguides; (<b>e</b>–<b>f</b>) FESEM images of the trapezoidal-shaped silicon waveguides fabricated on the (100) SOI wafer by using a KOH etching solution.</p>
Full article ">Figure 3
<p>The normalized electric field distributions of the WPP modes on the cross–section of the wedge SPP waveguides for various sidewall corners: (<b>a</b>) α = 35.26°, (<b>b</b>) 45°, (<b>c</b>) 54.7° and(<b>d</b>) 90°; and (<b>e</b>) the distributions of the light intensity along the AA’ cut line.</p>
Full article ">Figure 4
<p>The electric field distribution of the fundamental propagation mode on the cross–section of a trapezoidal-shaped SPP waveguide: (<b>a</b>) <span class="html-italic">E<sub>y</sub></span> electric field component and (<b>b</b>) <span class="html-italic">E<sub>z</sub></span> electric field component.</p>
Full article ">Figure 5
<p>The propagation characteristics of the WPP mode at various sidewall corners: (<b>a</b>) effective mode index (<span class="html-italic">n<sub>eff</sub></span>), (<b>b</b>) effective mode area (<span class="html-italic">A<sub>eff</sub></span>), (<b>c</b>) propagation length (<span class="html-italic">L</span>) and (<b>d</b>) the figure of merit (<span class="html-italic">FoM</span>).</p>
Full article ">Figure 6
<p>The propagation characteristics of the WPP mode depending on the width of the top surface of the wedge SPP waveguide, for four typical values of the sidewall corner (α = 35.26°, 45<sup>o</sup>, 54.7° and 90°): (<b>a</b>) effective mode index (<span class="html-italic">n<sub>eff</sub></span>), (<b>b</b>) effective ode area (<span class="html-italic">A<sub>eff</sub></span>), (<b>c</b>) propagation length (<span class="html-italic">L</span>) and (<b>d</b>) the figure of merit (<span class="html-italic">FoM</span>).</p>
Full article ">Figure 7
<p>The propagation characteristics of the wedge-shaped SPP waveguide structure as a function of the height (<b>α</b> = 54.74°, <span class="html-italic">w</span> = 0 nm, 300 nm, and 1000 nm): (<b>a</b>) effective mode index (<span class="html-italic">n<sub>eff</sub></span>), (<b>b</b>) effective mode area (<span class="html-italic">A<sub>eff</sub></span>), (<b>c</b>) propagation length (<span class="html-italic">L</span>) and (<b>d</b>) the figure of merit (<span class="html-italic">FoM</span>).</p>
Full article ">Figure 8
<p>The influence of the refractive index of the dielectric medium on the propagation characteristics: (<b>a</b>) effective mode index (<span class="html-italic">n<sub>eff</sub></span>), (<b>b</b>) effective mode area (<span class="html-italic">A<sub>eff</sub></span>), (<b>c</b>) propagation length (<span class="html-italic">L</span>) and (<b>d</b>) the figure of merit (<span class="html-italic">FoM</span>).</p>
Full article ">
9 pages, 2147 KiB  
Article
Temporal Distribution Measurement of the Parametric Spectral Gain in a Photonic Crystal Fiber Pumped by a Chirped Pulse
by Coralie Fourcade-Dutin, Antonio Imperio, Romain Dauliat, Raphael Jamier, Hector Muñoz-Marco, Pere Pérez-Millán, Hervé Maillotte, Philippe Roy and Damien Bigourd
Photonics 2019, 6(1), 20; https://doi.org/10.3390/photonics6010020 - 26 Feb 2019
Cited by 8 | Viewed by 3052
Abstract
The temporal distribution of the spectral parametric gain was experimentally investigated when a chirped pump pulse was injected into a photonic crystal fiber. A pump-probe experiment was developed and the important characteristics were measured as the chirp of the pump, the signal pulse, [...] Read more.
The temporal distribution of the spectral parametric gain was experimentally investigated when a chirped pump pulse was injected into a photonic crystal fiber. A pump-probe experiment was developed and the important characteristics were measured as the chirp of the pump, the signal pulse, and the gain of the parametric amplifier. We highlight that the amplified spectrum depends strongly on the instantaneous pump wavelength and that the temporal evolution of the wavelength at maximum gain is not monotonic. This behavior is significantly different from the case in which the chirped pump has a constant peak power. This measurement will be very important to efficiently include parametric amplifiers in laser systems delivering ultra-short pulses. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Second order dispersion term as a function of the wavelength. Inset. Input facet of the PCF measured with a scanning electron microscope. (<b>b</b>) Normalized pump spectrum of a chirped pulse with a Gaussian shape (black line) or with the experimental profile (red line).</p>
Full article ">Figure 2
<p>Phase matching as a function of the pump wavelength for a chirped pulse with a Gaussian spectrum (solid red and blue lines) or with the experimental profile (dashed red and blue lines). The maximum power was 600 W (red lines) or 200 W (blue lines). The magenta and cyan lines correspond to the phase matching for a constant peak power of 600 W or 200 W, respectively.</p>
Full article ">Figure 3
<p>Experimental set-up. VBG, volume Bragg grating; AOM, acoustooptic modulator; YDFA, ytterbium doped fiber amplifier; OSA, optical spectrum analyzer; PCF, photonic crystal fiber; ANDI, all normal dispersion fiber; BS, beam splitter; M, mirror; P, polarizer; DM, dichroic mirror; FM, flipped mirror.</p>
Full article ">Figure 4
<p>(<b>a</b>) Fluorescence spectrum when the PCF is pumped by 31 mW (black line). Output spectrum when the signal is injected at null delay (blue line). (<b>b</b>) Spectrum of the continuum generated in the ANDI fiber (black line) and spectrum of the signal (red line). (<b>c</b>) Autocorrelation trace (red stars) and the Gaussian fit (solid black line) of the signal.</p>
Full article ">Figure 5
<p>(<b>a</b>) Measurement of the chirp rate by spectral interferometry and its linear fit leading to α = 6.2 ps/nm. The inset is an example of the interference pattern. (<b>b</b>) Temporal shape of the pump pulse.</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) Selection of amplified and idler spectra when the delay is tuned from −6.6 ps to +19.8 ps. (<b>c</b>,<b>d</b>) Variation of the signal and idler wavelengths at the maximum amplification as a function of the delay.</p>
Full article ">Figure 7
<p>Maximum gain value (red circles) and its corresponding signal wavelength (black circles) as a function of the delay.</p>
Full article ">
14 pages, 3420 KiB  
Article
A Modified Design of a Hexagonal Circular Photonic Crystal Fiber with Large Negative Dispersion Properties and Ultrahigh Birefringence for Optical Broadband Communication
by Shovasis Kumar Biswas, Rishad Arfin, Ashfia Binte Habib, Syed Bin Amir, Zunayeed Bin Zahir, Mohammad Rezaul Islam and Md. Shahriar Hussain
Photonics 2019, 6(1), 19; https://doi.org/10.3390/photonics6010019 - 25 Feb 2019
Cited by 27 | Viewed by 5553
Abstract
In this paper, we propose a modified design of a hexagonal circular photonic crystal fiber (HC-PCF) which obtains a large negative dispersion and ultrahigh birefringence simultaneously. The optical properties of the proposed HC-PCF were investigated using the finite element method (FEM) incorporated with [...] Read more.
In this paper, we propose a modified design of a hexagonal circular photonic crystal fiber (HC-PCF) which obtains a large negative dispersion and ultrahigh birefringence simultaneously. The optical properties of the proposed HC-PCF were investigated using the finite element method (FEM) incorporated with a circular perfectly matched layer at the boundary. The simulation results showed large negative dispersion of −1044 ps/nm.km and ultrahigh birefringence of 4.321 × 10−2 at the operating wavelength of 1550 nm for the optimum geometrical parameters. Our proposed HC-PCF exhibited the desirable optical properties without non-circular air holes in the core and cladding region which facilitates the fabrication process. The large negative dispersion of the proposed microstructure over the wide spectral range, i.e., 1350 nm to 1600 nm, and high birefringence make it a suitable candidate for high-speed optical broadband communication and different sensing applications. Full article
(This article belongs to the Special Issue Lightwave Communications and Optical Networks)
Show Figures

Figure 1

Figure 1
<p>Cross-sectional view of the proposed hexagonal circular photonic crystal fiber (HC-PCF).</p>
Full article ">Figure 2
<p>(<b>a</b>) Electric field distribution of the fundamental mode at wavelength of <math display="inline"><semantics> <mrow> <mn>1550</mn> </mrow> </semantics></math> nm for <math display="inline"><semantics> <mi>x</mi> </semantics></math> polarization; (<b>b</b>) electric field distribution of the fundamental mode at wavelength of <math display="inline"><semantics> <mrow> <mn>1550</mn> </mrow> </semantics></math> nm for <math display="inline"><semantics> <mi>y</mi> </semantics></math> polarization; (<b>c</b>) convergence plot of the numerical simulation.</p>
Full article ">Figure 3
<p>Wavelength-dependent dispersion coefficient of the proposed HC-PCF for both <math display="inline"><semantics> <mi>x</mi> </semantics></math> and <math display="inline"><semantics> <mi>y</mi> </semantics></math> polarization for optimum geometrical parameters: <math display="inline"><semantics> <mrow> <mo>Λ</mo> <mo>=</mo> <mn>0.76</mn> <mtext> </mtext> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> <mo>,</mo> <mtext> </mtext> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> <mo>=</mo> <mn>0.97</mn> <mo>,</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> <mo>=</mo> <mn>0.45.</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Dispersion properties due to (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = 0.97, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = 0.45, and <math display="inline"><semantics> <mrow> <mo>Λ</mo> <mo>=</mo> <mrow> <mo>(</mo> <mrow> <mn>0.76</mn> <mtext> </mtext> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> <mo>,</mo> <mtext> </mtext> <mn>0.78</mn> <mtext> </mtext> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> <mo>,</mo> <mtext> </mtext> <mn>0.80</mn> <mtext> </mtext> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mo>Λ</mo> </semantics></math> = 0.76 μm, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math>= 0.45, and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mn>0.93</mn> <mtext> </mtext> <mn>0.95</mn> <mtext> </mtext> <mn>0.97</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mo>Λ</mo> </semantics></math>= 0.76 μm, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = 0.97, and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math>= <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mn>0.40</mn> <mtext> </mtext> <mn>0.43</mn> <mtext> </mtext> <mn>0.45</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> of the proposed HC-PCF.</p>
Full article ">Figure 4 Cont.
<p>Dispersion properties due to (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = 0.97, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = 0.45, and <math display="inline"><semantics> <mrow> <mo>Λ</mo> <mo>=</mo> <mrow> <mo>(</mo> <mrow> <mn>0.76</mn> <mtext> </mtext> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> <mo>,</mo> <mtext> </mtext> <mn>0.78</mn> <mtext> </mtext> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> <mo>,</mo> <mtext> </mtext> <mn>0.80</mn> <mtext> </mtext> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mo>Λ</mo> </semantics></math> = 0.76 μm, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math>= 0.45, and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mn>0.93</mn> <mtext> </mtext> <mn>0.95</mn> <mtext> </mtext> <mn>0.97</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mo>Λ</mo> </semantics></math>= 0.76 μm, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = 0.97, and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math>= <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mn>0.40</mn> <mtext> </mtext> <mn>0.43</mn> <mtext> </mtext> <mn>0.45</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> of the proposed HC-PCF.</p>
Full article ">Figure 5
<p>Birefringence properties of HC-PCF with (<b>a</b>) Λ = (0.76 μm, 0.78 μm, 0.80 μm) and fixed <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>0.97</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>0.45</mn> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mn>0.93</mn> <mtext> </mtext> <mn>0.95</mn> <mtext> </mtext> <mn>0.97</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> and fixed <math display="inline"><semantics> <mo>Λ</mo> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>0.76</mn> </mrow> </semantics></math> μm, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mo>=</mo> <mn>0.45</mn> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mo>Λ</mo> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>0.76</mn> </mrow> </semantics></math> μm, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>0.97</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mn>0.40</mn> <mtext> </mtext> <mn>0.43</mn> <mtext> </mtext> <mn>0.45</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Non-linear characteristics of the proposed HC-PCF for (<b>a</b>) both <math display="inline"><semantics> <mi>x</mi> </semantics></math> and <math display="inline"><semantics> <mi>y</mi> </semantics></math> polarization modes; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>0.97</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>0.45</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mo>Λ</mo> </mrow> </semantics></math> = (<math display="inline"><semantics> <mrow> <mn>0.76</mn> </mrow> </semantics></math> μm, <math display="inline"><semantics> <mrow> <mn>0.78</mn> </mrow> </semantics></math> μm, <math display="inline"><semantics> <mrow> <mn>0.80</mn> </mrow> </semantics></math> μm); (<b>c</b>) <math display="inline"><semantics> <mo>Λ</mo> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>0.76</mn> </mrow> </semantics></math> μm, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>0.45</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = (<math display="inline"><semantics> <mrow> <mn>0.93</mn> <mtext> </mtext> <mn>0.95</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mn>0.97</mn> </mrow> </semantics></math>) (<b>d</b>) <math display="inline"><semantics> <mo>Λ</mo> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>0.76</mn> </mrow> </semantics></math> μm, <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>0.97</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math> = (<math display="inline"><semantics> <mrow> <mn>0.40</mn> <mtext> </mtext> <mn>0.43</mn> <mtext> </mtext> <mn>0.45</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 7
<p>(<b>a</b>) Numerical aperture for two orthogonal polarization modes and (<b>b</b>) confinement loss of the proposed HC-PCF for optimum geometrical parameters.</p>
Full article ">Figure 8
<p>Effect on the (<b>a</b>) dispersion properties and (<b>b</b>) birefringence for <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>1</mn> <mo>%</mo> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>2</mn> <mo>%</mo> </mrow> </semantics></math> variation in the pitch.</p>
Full article ">Figure 9
<p>Effect on the (<b>a</b>) dispersion properties and (<b>b</b>) birefringence properties of the proposed HC-PCF for <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>1</mn> <mo>%</mo> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>2</mn> <mo>%</mo> </mrow> </semantics></math> variation in <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>Effect on the (<b>a</b>) dispersion properties and (<b>b</b>) birefringence properties of the proposed HC-PCF for <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>1</mn> <mo>%</mo> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>2</mn> <mo>%</mo> </mrow> </semantics></math> variation in <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>2</mn> </msub> <mo>/</mo> <mo>Λ</mo> </mrow> </semantics></math>.</p>
Full article ">
14 pages, 15136 KiB  
Article
Subwavelength Hexahedral Plasmonic Scatterers: History, Symmetries, and Resonant Characteristics
by Dimitrios Tzarouchis, Pasi Ylä-Oijala and Ari Sihvola
Photonics 2019, 6(1), 18; https://doi.org/10.3390/photonics6010018 - 25 Feb 2019
Cited by 2 | Viewed by 3277
Abstract
In this work, we investigate the resonant characteristics of hexahedral (cubical) inclusions at the plasmonic domain. After an introduction to the notion of superquadric surfaces, i.e., surfaces that model various versions of a rounded cube, we present the main resonant spectrum and the [...] Read more.
In this work, we investigate the resonant characteristics of hexahedral (cubical) inclusions at the plasmonic domain. After an introduction to the notion of superquadric surfaces, i.e., surfaces that model various versions of a rounded cube, we present the main resonant spectrum and the surface distributions for two particular cases of a smooth and a sharp cube in the plasmonic domain. We present a historical comparative overview of the main contributions available since the 1970s. A new categorization scheme of the resonances of a cube is introduced, based on symmetry considerations. The obtained results are compared against several recent works, exposing that the higher-order modes are extremely susceptible to both the choice of sharpness of the cube and the modeling mesh. This work can be readily used as a reference for both historical and contemporary studies of the plasmonic aspects of a cube. Full article
Show Figures

Figure 1

Figure 1
<p>The transformations of a superquadric sphere for increasing values of the rounding factor <span class="html-italic">p</span>. The bottom left and bottom right cases depict the studies cases used in this work: a “smooth” <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>15</mn> </mrow> </semantics></math> and a “sharp” <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math> cube. The figure indicates a constant <span class="html-italic">z</span>-directed electrostatic excitation <math display="inline"><semantics> <mrow> <mi mathvariant="bold">E</mi> <mo>=</mo> <msub> <mi>E</mi> <mn>0</mn> </msub> <msub> <mi mathvariant="bold">u</mi> <mi>z</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>The absolute value of the polarizability at the <math display="inline"><semantics> <mrow> <mi>ε</mi> <mo>∈</mo> <mo>[</mo> <mo>−</mo> <mn>5</mn> <mo>,</mo> <mn>0</mn> <mo>]</mo> </mrow> </semantics></math> domain of a smooth cube with <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>15</mn> </mrow> </semantics></math>. The color figures at the top color depict the surface potential of the six most pronounced resonances <math display="inline"><semantics> <msub> <mi>C</mi> <mn>1</mn> </msub> </semantics></math>–<math display="inline"><semantics> <msub> <mi>C</mi> <mn>6</mn> </msub> </semantics></math>, for the regular mesh, while at the bottom the same surface potential are shown for the refined mesh. Blue and red colors indicate negative and positive values of the potential. The inset figure illustrates the used meshes, i.e., regular and refined ones.</p>
Full article ">Figure 3
<p>As in <a href="#photonics-06-00018-f002" class="html-fig">Figure 2</a>, the absolute value of the polarizability at the <math display="inline"><semantics> <mrow> <mi>ε</mi> <mo>∈</mo> <mo>[</mo> <mo>−</mo> <mn>5</mn> <mo>,</mo> <mn>0</mn> <mo>]</mo> </mrow> </semantics></math> domain for the sharp cube (<math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math>). The color figures at the top depict the surface potential of the six most pronounced resonances <math display="inline"><semantics> <msub> <mi>C</mi> <mn>1</mn> </msub> </semantics></math>–<math display="inline"><semantics> <msub> <mi>C</mi> <mn>6</mn> </msub> </semantics></math>, extracted with a regular mesh, while the bottom figures are extracted with a refined mesh. The <math display="inline"><semantics> <msub> <mi>C</mi> <mn>4</mn> </msub> </semantics></math> resonance is actually different between the two mesh setups (inset sketches top and bottom). Note also the disagreement of the <math display="inline"><semantics> <msub> <mi>C</mi> <mn>3</mn> </msub> </semantics></math> resonance between the smooth case (<a href="#photonics-06-00018-f002" class="html-fig">Figure 2</a>) and the sharp case (here) [<a href="#B15-photonics-06-00018" class="html-bibr">15</a>]. The observed inconsistencies provide evidence that the higher-order modes residing at <math display="inline"><semantics> <mrow> <mi>ε</mi> <mo>∈</mo> <mo>(</mo> <mo>−</mo> <mn>3</mn> <mo>,</mo> <mo>−</mo> <mn>1</mn> <mo>)</mo> </mrow> </semantics></math> are sensitive to both the mesh and the sharpness of the cube.</p>
Full article ">Figure 4
<p>A sketch describing the proposed categorization scheme: vertex-edge vectors <math display="inline"><semantics> <msub> <mi mathvariant="bold">V</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi mathvariant="bold">V</mi> <mn>2</mn> </msub> </semantics></math> are denoted with red, and face (hedral) vectors <math display="inline"><semantics> <msub> <mi mathvariant="bold">F</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi mathvariant="bold">F</mi> <mn>2</mn> </msub> </semantics></math> with green. The blue vector indicates the axis of rotational symmetry due to the used excitation (<math display="inline"><semantics> <msub> <mi mathvariant="bold">u</mi> <mi>z</mi> </msub> </semantics></math>-polarized electrostatic field). A mode is characterized by four numbers, denoting the number of nodes (sign changes of the surface potential distribution) along the four vectors <math display="inline"><semantics> <mrow> <mo>(</mo> <msub> <mi>V</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>V</mi> <mn>2</mn> </msub> <mo>,</mo> <msub> <mi>F</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>F</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> </semantics></math>. The combination of the four numbers and the particular symmetries of cube, allow the recreation of every mode. Please note that in the case of a sphere all <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="bold">V</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi mathvariant="bold">F</mi> <mn>1</mn> </msub> <mo>,</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <msub> <mi mathvariant="bold">F</mi> <mn>2</mn> </msub> </semantics></math> are parallel (or antiparallel) representing the elevation vector <math display="inline"><semantics> <msub> <mi mathvariant="bold">u</mi> <mi>θ</mi> </msub> </semantics></math>, while <math display="inline"><semantics> <msub> <mi mathvariant="bold">V</mi> <mn>2</mn> </msub> </semantics></math> becomes the azimuthal vector <math display="inline"><semantics> <msub> <mi mathvariant="bold">u</mi> <mi>ϕ</mi> </msub> </semantics></math>.</p>
Full article ">Figure 5
<p>The six main resonances and the three particular clusters for the case of a smooth cube with <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>15</mn> </mrow> </semantics></math> and a regular mesh. A total of 22 resonances and their surface potential distribution is presented. Above each resonance the categorization number is given, described in <a href="#photonics-06-00018-f004" class="html-fig">Figure 4</a>. Please note that all resonances in clusters are more sensitive to modeling mesh density, than the main six resonances <math display="inline"><semantics> <mrow> <msub> <mi>C</mi> <mn>1</mn> </msub> <mo>−</mo> <msub> <mi>C</mi> <mn>6</mn> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>The six main resonances and the three particular clusters for the case of a sharp cube with <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math> and a refined mesh. A total of 17 resonances and their surface potential distributions are presented. Above each resonance the categorization number is given, described in <a href="#photonics-06-00018-f004" class="html-fig">Figure 4</a>. Please note that all resonances in clusters are more sensitive to modeling mesh density, than the main six resonances <math display="inline"><semantics> <mrow> <msub> <mi>C</mi> <mn>1</mn> </msub> <mo>−</mo> <msub> <mi>C</mi> <mn>6</mn> </msub> </mrow> </semantics></math>.</p>
Full article ">
10 pages, 2525 KiB  
Communication
A Novel Microtiter Plate Format High Power Open Source LED Array
by Heike Kagel, Hannes Jacobs, Frank F. Bier, Jörn Glökler and Marcus Frohme
Photonics 2019, 6(1), 17; https://doi.org/10.3390/photonics6010017 - 25 Feb 2019
Cited by 6 | Viewed by 4853
Abstract
Many photochemical or photobiological applications require the use of high power ultraviolet light sources, such as high-pressure mercury arc lamps. In addition, many photo-induced chemical, biochemical and biological applications require either a combinatorial setting or a parallel assay of multiple samples under the [...] Read more.
Many photochemical or photobiological applications require the use of high power ultraviolet light sources, such as high-pressure mercury arc lamps. In addition, many photo-induced chemical, biochemical and biological applications require either a combinatorial setting or a parallel assay of multiple samples under the same environmental conditions to ensure reproducibility. To achieve this, alternative, controllable light sources, such as ultraviolet light emitting diodes (UV LEDs) with high power and spatial control are required. Preferably, LEDs are arranged in a suitable standardized 96-well microtiter plate format. We designed such an array and established the methods required for heat management and enabling stable, controllable illumination over time. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic overview of Setup 1. Left side CAD drawn design (<b>a</b>), (<b>b</b>) photograph from top view, (<b>c</b>) setup 1 with mounted microtiter plate. Red rectangle in a) indicates the row were IR images were recorded. The setup consisted of 3 × 6 LEDs. Each row was powered by a LDD-700 L constant-current source, and could be individually controlled by a microcontroller.</p>
Full article ">Figure 2
<p>Schematic overview of setup 2. (<b>a</b>) shows the CAD design, (<b>b</b>) a photograph from setup 2. Red rectangle in a) indicates the row were IR images were recorded.</p>
Full article ">Figure 3
<p>Schematic overview of setup 3, (<b>a</b>) shows the CAD design, (<b>b</b>) a photograph from setup 2. Red rectangle in a) indicates the row were IR images were recorded.</p>
Full article ">Figure 4
<p>LED emission intensity measurement setup.</p>
Full article ">Figure 5
<p>Casing Temperature of the three different setups when using LEDs in CW mode after 5 min with different fixing methods (<b>a</b>) with thermal conductive tape in setup 1 (<b>b</b>) left thermal conductive tape, right thermal conductive adhesive paste in setup 2 (<b>c</b>) thermal conductive adhesive paste with radial fans on in setup 3.</p>
Full article ">Figure 6
<p>Light intensities recorded of (<b>a</b>) setup 1, (<b>b</b>) setup 2, (<b>c</b>) setup 3 without radial fans turned on and (<b>d</b>) setup 3 with radial fans turned on.</p>
Full article ">
13 pages, 379 KiB  
Article
Optical Space Switches in Data Centers: Issues with Transport Protocols
by Mamun Abu-Tair, Philip Perry, Philip Morrow, Sally McClean, Bryan Scotney, Gerard Parr and Md Israfil Biswas
Photonics 2019, 6(1), 16; https://doi.org/10.3390/photonics6010016 - 22 Feb 2019
Cited by 1 | Viewed by 3209
Abstract
A number of new architectures for data centre networks employing reconfigurable, SDN controlled, all-optical networks have been reported in recent years. In most cases, additional capacity was added to the system which unsurprisingly improved performance. In this study, a generalised network model that [...] Read more.
A number of new architectures for data centre networks employing reconfigurable, SDN controlled, all-optical networks have been reported in recent years. In most cases, additional capacity was added to the system which unsurprisingly improved performance. In this study, a generalised network model that emulates the behaviour of these types of network was developed but where the total capacity is maintained constant so that system behaviour can be understood. An extensive emulated study is presented which indicates that the reconfiguration of such a network can have a detrimental impact on Transmission Control Protocol (TCP) congestion control mechanisms that can degrade the performance of the system. A number of simple scheduling mechanisms were investigated and the results show that an on-demand scheduling mechanism could deliver a throughput increase of more than ∼50% without any increase in total installed network capacity. These results, therefore, indicate the need to link the network resource management with new datacentre network architectures. Full article
(This article belongs to the Special Issue Lightwave Communications and Optical Networks)
Show Figures

Figure 1

Figure 1
<p>Direct-Optical Connection part of the Data Centre Network (DCN).</p>
Full article ">Figure 2
<p>Schematic of emulated network used in the study showing Hosts/Switches connectivity. All Links have a 1 Gbps bandwidth limit.</p>
Full article ">Figure 3
<p>Average throughput of 24 VM transfers against different switching epochs.</p>
Full article ">Figure 4
<p>Throughput for elephant and mice VMs in the 24 (<b>a</b>), 48 (<b>b</b>) and 72 (<b>c</b>) VM transfer scenarios. To the right of each boxplot, we show the mean and a whisker marking the 95th and 99th percentiles.</p>
Full article ">Figure 5
<p>Completion time for elephant and mice VMs in the 24 (<b>a</b>), 48 (<b>b</b>) and 72 (<b>c</b>) VM transfer scenarios. To the right of each boxplot, we show the mean and a whisker marking the 95th and 99th percentiles.</p>
Full article ">Figure 6
<p>Throughput for elephant and mice VMs in the 24 (<b>a</b>), 48 (<b>b</b>) and 72 (<b>c</b>) VM transfer scenarios. To the right of each boxplot, we show the mean and a whisker marking the 95th and 99th percentiles.</p>
Full article ">Figure 7
<p>Completion time for elephant and mice VMs in the 24 (<b>a</b>), 48 (<b>b</b>) and 72 (<b>c</b>) VM transfer scenarios. To the right of each boxplot, we show the mean and a whisker marking the 95th and 99th percentiles.</p>
Full article ">Figure 8
<p>Throughput for elephant and mice VMs in the 24 (<b>a</b>), 48 (<b>b</b>) and 72 (<b>c</b>) VM transfer scenarios. To the right of each boxplot, we show the mean and a whisker marking the 95th and 99th percentiles.</p>
Full article ">Figure 9
<p>Completion time for elephant and mice VMs in the 24 (<b>a</b>), 48 (<b>b</b>) and 72 (<b>c</b>) VM transfer scenarios. To the right of each boxplot, we show the mean and a whisker marking the 95th and 99th percentiles.</p>
Full article ">
13 pages, 1097 KiB  
Article
International System of Units (SI) Traceable Noise-Equivalent Power and Responsivity Characterization of Continuous Wave ErAs:InGaAs Photoconductive Terahertz Detectors
by Anuar de Jesus Fernandez Olvera, Axel Roggenbuck, Katja Dutzi, Nico Vieweg, Hong Lu, Arthur C. Gossard and Sascha Preu
Photonics 2019, 6(1), 15; https://doi.org/10.3390/photonics6010015 - 13 Feb 2019
Cited by 16 | Viewed by 4766
Abstract
A theoretical model for the responsivity and noise-equivalent power (NEP) of photoconductive antennas (PCAs) as coherent, homodyne THz detectors is presented. The model is validated by comparison to experimental values obtained for two ErAs:InGaAs PCAs. The responsivity and NEP were obtained from the [...] Read more.
A theoretical model for the responsivity and noise-equivalent power (NEP) of photoconductive antennas (PCAs) as coherent, homodyne THz detectors is presented. The model is validated by comparison to experimental values obtained for two ErAs:InGaAs PCAs. The responsivity and NEP were obtained from the measured rectified current, the current noise floor in the PCAs, and the incoming THz power for the same conditions. Since the THz power measurements are performed with a pyroelectric detector calibrated by the National Metrology Institute of Germany (PTB), the experimentally obtained values are directly traceable to the International System of Units (SI) for the described conditions. The agreement between the presented model and the experimental results is excellent using only one fitting parameter. A very low NEP of 1.8 fW/Hz at 188.8 GHz is obtained at room temperature. Full article
(This article belongs to the Special Issue Terahertz Photonics)
Show Figures

Figure 1

Figure 1
<p>Geometric layout of the finger electrode (black) structure on the photoconductive material (blue) with geometric variables as defined in the text: (<b>a</b>) Top view (<b>b</b>) Cross-sectional view (red lines depict the field lines across the gap).</p>
Full article ">Figure 2
<p>Equivalent circuit of the receiving antenna including the incoming THz wave with power <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mrow> <mi>T</mi> <mi>H</mi> <mi>z</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>f</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>, the antenna and the photoconductor. The value of <math display="inline"><semantics> <mrow> <msub> <mi>U</mi> <mrow> <mi>T</mi> <mi>H</mi> <mi>z</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>f</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> is the Thevenin equivalent open-circuit voltage source, as defined in [<a href="#B16-photonics-06-00015" class="html-bibr">16</a>].</p>
Full article ">Figure 3
<p>Experimental setup for the measurements. For THz power measurement, the detector was a pyroelectric device, calibrated by the PTB. For the THz current measurements, it was replaced by the PCAs.</p>
Full article ">Figure 4
<p>Squared detected THz current versus measured THz power for PCA B. The fitted responsivity <math display="inline"><semantics> <mrow> <mi>S</mi> <mo>(</mo> <mi>f</mi> <mo>)</mo> </mrow> </semantics></math> (shown in dashed lines) was equated to Equation (<a href="#FD10-photonics-06-00015" class="html-disp-formula">10</a>) in order to obtain the effective THz effiency <math display="inline"><semantics> <msubsup> <mi>η</mi> <mrow> <mi>T</mi> <mi>H</mi> <mi>z</mi> </mrow> <mrow> <mi>e</mi> <mi>f</mi> <mi>f</mi> </mrow> </msubsup> </semantics></math>. The minimum THz power that could be measured was limited by the sensitivity of the PTB calibrated detector (1 <math display="inline"><semantics> <mi>μ</mi> </semantics></math>W).</p>
Full article ">Figure 5
<p>Comparison between the calculated and experimentally obtained responsivities of (<b>a</b>) PCA A (<b>b</b>) PCA B. <math display="inline"><semantics> <msubsup> <mi>η</mi> <mrow> <mi>T</mi> <mi>H</mi> <mi>z</mi> </mrow> <mrow> <mi>e</mi> <mi>f</mi> <mi>f</mi> </mrow> </msubsup> </semantics></math> is assumed frequency-independent.</p>
Full article ">Figure 6
<p>Comparison between the calculated (<math display="inline"><semantics> <msubsup> <mi>η</mi> <mrow> <mi>T</mi> <mi>H</mi> <mi>z</mi> </mrow> <mrow> <mi>e</mi> <mi>f</mi> <mi>f</mi> </mrow> </msubsup> </semantics></math> = 0.012, <math display="inline"><semantics> <mrow> <msub> <mi>I</mi> <mrow> <mi>n</mi> <mi>o</mi> <mi>i</mi> <mi>s</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>I</mi> <mrow> <mi>t</mi> <mi>h</mi> <mi>e</mi> <mi>r</mi> <mi>m</mi> </mrow> </msub> </mrow> </semantics></math>), the ideal (with <math display="inline"><semantics> <msubsup> <mi>η</mi> <mrow> <mi>T</mi> <mi>H</mi> <mi>z</mi> </mrow> <mrow> <mi>e</mi> <mi>f</mi> <mi>f</mi> </mrow> </msubsup> </semantics></math> = 1, <math display="inline"><semantics> <mrow> <msub> <mi>I</mi> <mrow> <mi>n</mi> <mi>o</mi> <mi>i</mi> <mi>s</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>I</mi> <mrow> <mi>t</mi> <mi>h</mi> <mi>e</mi> <mi>r</mi> <mi>m</mi> </mrow> </msub> </mrow> </semantics></math>), and the experimentally obtained NEPs for PCA A and B. The figure also shows the ideal performance of PCA C.</p>
Full article ">Figure A1
<p>Measured THz powers emitted by the pin diode for the characterization of PCA A and B.</p>
Full article ">Figure A2
<p>Comparison between the experimentally obtained and the calculated rectified current of PCA A and B.</p>
Full article ">
11 pages, 892 KiB  
Article
Grating Lobes in Higher-Order Correlation Functions of Arrays of Quantum Emitters: Directional Photon Bunching Versus Correlated Directions
by Iñigo Liberal, Iñigo Ederra and Richard W. Ziolkowski
Photonics 2019, 6(1), 14; https://doi.org/10.3390/photonics6010014 - 12 Feb 2019
Cited by 9 | Viewed by 3607
Abstract
Recent advances in nanofabrication and optical manipulation techniques are making it possible to build arrays of quantum emitters with accurate control over the locations of their individual elements. In analogy with classical antenna arrays, this poses new opportunities for tailoring quantum interference effects [...] Read more.
Recent advances in nanofabrication and optical manipulation techniques are making it possible to build arrays of quantum emitters with accurate control over the locations of their individual elements. In analogy with classical antenna arrays, this poses new opportunities for tailoring quantum interference effects by designing the geometry of the array. Here, we investigate the N th -order directional correlation function of the photons emitted by an array of N initially-excited identical quantum emitters, addressing the impact of the appearance of grating lobes. Our analysis reveals that the absence of directivity in the first-order correlation function is contrasted by an enhanced directivity in the N th -order one. This suggests that the emitted light consists of a superposition of directionally entangled photon bunches. Moreover, the photon correlation landscape changes radically with the appearance of grating lobes. In fact, the photons no longer tend to be bunched along the same direction; rather, they are distributed in a set of correlated directions with equal probability. These results clarify basic aspects of light emission from ensembles of quantum emitters. Furthermore, they may find applications in the design of nonclassical light sources. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Sketch of the geometry. An ensemble of <span class="html-italic">N</span> quantum emitters are located at positions <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="bold">r</mi> <mn>1</mn> </msub> <mo>,</mo> <mo>…</mo> <mo>,</mo> <msub> <mi mathvariant="bold">r</mi> <mi>N</mi> </msub> </mrow> </semantics></math>. They are modelled as two-level systems <math display="inline"><semantics> <mfenced separators="" open="{" close="}"> <mfenced open="|" close="&#x232A;"> <mi>e</mi> </mfenced> <mo>,</mo> <mfenced open="|" close="&#x232A;"> <mi>g</mi> </mfenced> </mfenced> </semantics></math>. All of these emitters are initially excited <math display="inline"><semantics> <mrow> <mfenced open="|" close="&#x232A;"> <mi>ψ</mi> </mfenced> <mo>=</mo> <mfenced separators="" open="|" close="&#x232A;"> <msub> <mi>e</mi> <mn>1</mn> </msub> <mo>⋯</mo> <msub> <mi>e</mi> <mi>N</mi> </msub> </mfenced> </mrow> </semantics></math> and decay by emitting photons. The photon emission occurs in correlated directions that depend on the geometry of the array. (<b>b</b>) Conceptual sketch of directional photon bunching. The average number of photons measured in a given direction has an isotropic emission pattern (solid red line), but each array decay process exhibits bunching of the photons along specific directions (dashed lines).</p>
Full article ">Figure 2
<p>(<b>a</b>) Sketch of the geometry. A linear array of <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> uniformly spaced quantum emitters with position vectors <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="bold">r</mi> <mi>n</mi> </msub> <mo>=</mo> <mi>n</mi> <mi>d</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>,</mo> <mn>3</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.5</mn> <msub> <mi>λ</mi> <mn>0</mn> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <msub> <mi>λ</mi> <mn>0</mn> </msub> </semantics></math> being the wavelength at their transition frequency, is oriented vertically along the Z-axis. (<b>b</b>) Sketch of three measurement outcomes with the same probability. Top-left: Three photons measured along the same direction <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mo>=</mo> <msub> <mi>θ</mi> <mn>3</mn> </msub> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>3</mn> </mrow> </semantics></math>. Top-right: Two photons measured along this direction: <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mo>=</mo> <msub> <mi>θ</mi> <mn>3</mn> </msub> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>3</mn> </mrow> </semantics></math>, and the remaining photon measured along the first grating lobe direction <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mo>=</mo> <mi>acos</mi> <mrow> <mo stretchy="false">(</mo> <mi>cos</mi> <mfrac> <mi>π</mi> <mn>3</mn> </mfrac> <mo>−</mo> <mfrac> <mrow> <mn>2</mn> <mi>π</mi> </mrow> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>d</mi> </mrow> </mfrac> <mo>≃</mo> <mn>0.55</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math>. Bottom-center: Two photons measured along the direction <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mo>=</mo> <msub> <mi>θ</mi> <mn>3</mn> </msub> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>3</mn> </mrow> </semantics></math>, and the remaining photon measured along the second grating lobe direction, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mo>=</mo> <mi>acos</mi> <mrow> <mo stretchy="false">(</mo> <mi>cos</mi> <mfrac> <mi>π</mi> <mn>3</mn> </mfrac> <mo>−</mo> <mfrac> <mrow> <mn>4</mn> <mi>π</mi> </mrow> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>d</mi> </mrow> </mfrac> <mo>≃</mo> <mn>0.81</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math>. For reference, the red line indicates the <math display="inline"><semantics> <msup> <mi>N</mi> <mi>th</mi> </msup> </semantics></math>-order array factor <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mn>3</mn> </msub> <mrow> <mo stretchy="false">(</mo> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mo>,</mo> <msub> <mi>θ</mi> <mn>3</mn> </msub> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> as a function of <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>1</mn> </msub> </semantics></math> and evaluated at <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mo>=</mo> <msub> <mi>θ</mi> <mn>3</mn> </msub> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>3</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Quantum array factor (normalized to its maximum value) for different observation angles and emitter separation distances for a vertical linear array of <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> quantum emitters uniformly spaced along the Z-axis. The transition dipole moment of the emitters is assumed for the convenience to be oriented along the Z-axis, <math display="inline"><semantics> <mrow> <mi mathvariant="bold">p</mi> <mo>=</mo> <msub> <mi mathvariant="bold">u</mi> <mi>z</mi> </msub> <mspace width="0.166667em"/> <mi>p</mi> </mrow> </semantics></math>, in order to make the quantum array factor depend only on the elevation angle <math display="inline"><semantics> <mi>θ</mi> </semantics></math>. The normalized <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mi>N</mi> <mo>=</mo> <mn>3</mn> <mo stretchy="false">)</mo> </mrow> </semantics></math>-order quantum array factor, <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mn>3</mn> </msub> <mrow> <mo stretchy="false">(</mo> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mo>,</mo> <msub> <mi>θ</mi> <mn>3</mn> </msub> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math>, is explicitly shown as a function of <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>2</mn> </msub> </semantics></math> for different values of <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>3</mn> </msub> </semantics></math>: first column, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.5</mn> <mi>π</mi> </mrow> </semantics></math>; second column, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.3</mn> <mi>π</mi> </mrow> </semantics></math>; and third column, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.1</mn> <mi>π</mi> </mrow> </semantics></math>; and as a function of the emitter separation distance <span class="html-italic">d</span>: first row, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>0.1</mn> <msub> <mi>λ</mi> <mn>0</mn> </msub> </mrow> </semantics></math>; second row, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.0</mn> <msub> <mi>λ</mi> <mn>0</mn> </msub> </mrow> </semantics></math>; and third row, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>2.5</mn> <msub> <mi>λ</mi> <mn>0</mn> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <msub> <mi>λ</mi> <mn>0</mn> </msub> </semantics></math> being the wavelength at the transition frequency of the quantum emitters.</p>
Full article ">
40 pages, 14703 KiB  
Review
Recent Trends and Advances of Silicon-Based Integrated Microwave Photonics
by Reza Maram, Saket Kaushal, José Azaña and Lawrence R Chen
Photonics 2019, 6(1), 13; https://doi.org/10.3390/photonics6010013 - 30 Jan 2019
Cited by 50 | Viewed by 9664
Abstract
Multitude applications of photonic devices and technologies for the generation and manipulation of arbitrary and random microwave waveforms, at unprecedented processing speeds, have been proposed in the literature over the past three decades. This class of photonic applications for microwave engineering is known [...] Read more.
Multitude applications of photonic devices and technologies for the generation and manipulation of arbitrary and random microwave waveforms, at unprecedented processing speeds, have been proposed in the literature over the past three decades. This class of photonic applications for microwave engineering is known as microwave photonics (MWP). The vast capabilities of MWP have allowed the realization of key functionalities which are either highly complex or simply not possible in the microwave domain alone. Recently, this growing field has adopted the integrated photonics technologies to develop microwave photonic systems with enhanced robustness as well as with a significant reduction of size, cost, weight, and power consumption. In particular, silicon photonics technology is of great interest for this aim as it offers outstanding possibilities for integration of highly-complex active and passive photonic devices, permitting monolithic integration of MWP with high-speed silicon electronics. In this article, we present a review of recent work on MWP functions developed on the silicon platform. We particularly focus on newly reported designs for signal modulation, arbitrary waveform generation, filtering, true-time delay, phase shifting, beam steering, and frequency measurement. Full article
Show Figures

Figure 1

Figure 1
<p>Architecture of a generic microwave photonic system. AWG: arbitrary waveform generation; MWP: microwave photonics; RF: radio-frequency.</p>
Full article ">Figure 2
<p>(<b>a</b>) Top-view scanning electron microscope image of the fabricated silicon-on-insulator microring coupled-resonator optical waveguides (CROW) filter reported in [<a href="#B81-photonics-06-00013" class="html-bibr">81</a>]. (<b>b</b>) Measured and theoretical ((amplitude (left) and phase (right)) response of sixth-order CROW filter. (<b>c</b>) Optical spectrum of the generated optical single-sideband (OSSB) signal with lower sideband suppression (black line) and upper sideband suppression (red line) when modulated at 20 GHz (after passing through the optical filter). (<b>d</b>) RF responses of the optical dual sideband (ODSB) and OSSB system with (left) upper sideband suppression (red dashed line) (right) lower sideband suppression (black dashed line). Courtesy of Xiaoke Yi [<a href="#B81-photonics-06-00013" class="html-bibr">81</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic illustration of the photonic assisted microwave frequency-modulation system using an integrated silicon Mach Zehnder modulator (MZM) reported in [<a href="#B83-photonics-06-00013" class="html-bibr">83</a>]; PC: polarization controller; EDFA: erbium-doped fiber amplifier; VOA: variable optical attenuator; TLD: tunable laser diode; PD: photodetector. (<b>b,c</b>) From top to bottom: waveforms of the original 50-Mb/s baseband signal, original 1-GHz microwave carrier signal, and the output microwave amplitude-shift keying (ASK) signal respectively. Courtesy of Jian Wang [<a href="#B83-photonics-06-00013" class="html-bibr">83</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) Illustration of the principle for amplitude modulation of a monocycle pulse. (<b>b</b>) Principle for bi-phase modulation reported in [<a href="#B85-photonics-06-00013" class="html-bibr">85</a>]. (<b>c</b>) Electrical spectrum of the monocycle pulses with repetition rate of 625 MHz for both positive and negative polarity. (<b>d</b>) RF spectrum of the amplitude modulated monocycle pulses. Inset: waveform of the amplitude modulated pulses with a pattern of “01110” without zero padding. (<b>e</b>) RF spectrum of the bi-phase modulated monocycle pulses. Inset: waveform of the phase modulated pulses with a pattern of “01110” without zero padding. Courtesy of Hon Ki Tsang [<a href="#B85-photonics-06-00013" class="html-bibr">85</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) Schematic of the dual-phase-shifted waveguide Bragg grating (DPS-WBG) reported in [<a href="#B87-photonics-06-00013" class="html-bibr">87</a>]. (<b>b</b>) DPS-WBG magnitude response, (<b>c</b>) Operation as phase shifter (PS), and (<b>d</b>) true time delay (TTD). (<b>e</b>) Complex (magnitude and phase) RF response for different phase shift values. (<b>f</b>) Complex (magnitude and phase) RF response for different true time delay values. Courtesy of Maurizio Burla [<a href="#B87-photonics-06-00013" class="html-bibr">87</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic of a microwave photonic phase shifter using an SOI MRR (silicon-on-insulator microring resonator) reported in [<a href="#B34-photonics-06-00013" class="html-bibr">34</a>]. OSSB: optical single-sideband, PD: photo-detector. Inset: the amplitude and phase responses of the SOI-MRR. (<b>b</b>) The phase response of the three cascaded microring resonators (MRRs) showing a net redshift of the phase response with the increase of the pumping power from 0 (no pump) to 27 dBm. (<b>c</b>) Measured phase shifts at different pumping power levels. Courtesy of Jianping Yao [<a href="#B34-photonics-06-00013" class="html-bibr">34</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic and operation principle of the proposed photonic integrated phase shifter reported in [<a href="#B88-photonics-06-00013" class="html-bibr">88</a>]. LS: laser source; SSB EOM: single-sideband electro-optical modulator. In the scheme, <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mrow> <mi>R</mi> <mi>F</mi> </mrow> </msub> </mrow> </semantics></math> represents the RF signal carrier frequency, <math display="inline"><semantics> <mrow> <msub> <mi>ν</mi> <mi>c</mi> </msub> </mrow> </semantics></math> the optical carrier frequency, <math display="inline"><semantics> <mrow> <msub> <mi>ν</mi> <mrow> <mi>s</mi> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math> is the central optical frequency of the optical SSB modulated signal, <math display="inline"><semantics> <mi>ϕ</mi> </semantics></math> is the variable optical/RF phase shift. (<b>b</b>) Picture of the fabricated PIC; GC: grating coupler; MMI-OC: multi-mode interference optical coupler. Two PDs are connected to the two MMI outputs. (<b>c</b>) Magnitude (top) and phase (bottom) response of the microwave photonics-phase shifter MWP-PS, as a function of the vector network analyzer (VNA) output RF frequency, over a 10-to-16 GHz frequency span. (<b>d</b>) Magnitude (top) and phase (bottom) response over a 36-to-42 GHz frequency span. Courtesy of Paolo Ghelfi [<a href="#B88-photonics-06-00013" class="html-bibr">88</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) Perspective view of an electrically tunable silicon-based on-chip microdisk resonator (MDR) reported in [<a href="#B93-photonics-06-00013" class="html-bibr">93</a>]. (<b>b</b>) Top view of the proposed MDR. (<b>c</b>) Measured transmission spectrum of the fabricated MDR, and (<b>d</b>) zoom-in view of the resonance of the fabricated MDR at the wavelength of 1551.84 nm. (<b>e</b>) Measured group delay of the resonance WGM_0,38 in the fabricated MDR and (<b>f</b>) electrical tunability of the group delay at three different wavelengths. Courtesy of Jianping Yao [<a href="#B93-photonics-06-00013" class="html-bibr">93</a>].</p>
Full article ">Figure 9
<p>(<b>a</b>) The waveguide cross-section. (<b>b</b>) Parameters of the subwavelength grating (SWG) waveguide. (<b>c</b>) Schematic of the fabricated 4-tap optical true time delay line (OTTDL) structure based on SWG waveguides reported in [<a href="#B94-photonics-06-00013" class="html-bibr">94</a>]. (<b>d</b>) Measured power spectral response of the fabricated OTTDL. (<b>e</b>) Generated time-domain pulse train at the output of the fabricated OTTDL device in response to a single input optical pulse. (<b>f</b>) Array of 4 SWG waveguides with varying duty cycles from D<sub>1</sub> = 60% to D<sub>4</sub> = 30% in increments of 10%. (<b>g</b>) The measured RF phase shift when the modulated light is transmitted through different taps of the OTTDL. (<b>h</b>) The calculated probability distribution function (PDF) of the relative time delay given by the RF phase slope versus frequency shown in (<b>g</b>). The measured incremental time delay between the taps are 8.9 ps, 10.7 ps, and 7.9 ps, respectively.</p>
Full article ">Figure 10
<p>(<b>a</b>) Architecture of the continuously tunable optical delay line structure reported in [<a href="#B95-photonics-06-00013" class="html-bibr">95</a>]. (<b>b</b>) Cross-sectional structure of the ultra-thin waveguide with a titanium nitride (TiN) heater on top. (<b>c</b>) Simulated x-component of the electronic field distribution for the fundamental transverse electric (TE) mode. (<b>d</b>) Dual-ring slow-light delay line structure. (<b>e</b>) Evolution of the ring delay spectrum upon thermal tuning. (<b>f</b>) Optical pulse waveforms after passing through the switchable delay line. Black curves: reference pulses; blue curves: delayed pulses. (<b>g</b>) Optical pulse waveforms after passing through the longest optical path with delay fine tuning by the ring resonators. (<b>h</b>) Eye diagrams of a 30 Gb/s pseudo-random binary sequences (PRBS) signal after various delays; Gb/s: gigabits per second. Courtesy of Linjie Zhou [<a href="#B95-photonics-06-00013" class="html-bibr">95</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) Schematic configuration of the proposed 1D grating waveguide reported in [<a href="#B96-photonics-06-00013" class="html-bibr">96</a>]; (<b>b</b>) the electric field intensity distribution of the mode in the xz plane and (<b>c</b>) in the yz plane. (<b>d</b>) Wavelength continuously tunable beam steering system based on the 1D grating waveguide true time delay (TTD) lines. Amp: amplifier. Inset: schematic of the enlarged spiraled strip waveguide. (<b>e</b>) Delay time and beam steering angle versus wavelength for all four channels. The lengths of 1D grating waveguides are 0 mm, 1 mm, 2 mm, and 3 mm corresponding to channel 0, channel 1, channel 2, and channel 3. Courtesy of Jianyi Yang [<a href="#B96-photonics-06-00013" class="html-bibr">96</a>].</p>
Full article ">Figure 12
<p>(<b>a</b>) Schematic diagram for microwave frequency measurement reported in [<a href="#B102-photonics-06-00013" class="html-bibr">102</a>]. (<b>b</b>) and (<b>c</b>) Calculated microwave photonic filter (MPF) responses and amplitude comparison function (ACF), respectively. (<b>d</b>) Measured MPF responses for three different wavelengths of <math display="inline"><semantics> <mrow> <msub> <mi>λ</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>λ</mi> <mn>2</mn> </msub> <mo>,</mo> <msub> <mi>λ</mi> <mn>3</mn> </msub> </mrow> </semantics></math> (<b>e</b>) Theoretical and measured ACF<sub>12</sub> when the system is operating at <math display="inline"><semantics> <mrow> <msub> <mi>λ</mi> <mn>1</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>λ</mi> <mn>2</mn> </msub> </mrow> </semantics></math>, respectively. (<b>f</b>) Theoretical and measured ACF<sub>12</sub> when the system is operating at <math display="inline"><semantics> <mrow> <msub> <mi>λ</mi> <mn>2</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>λ</mi> <mn>3</mn> </msub> </mrow> </semantics></math>, respectively. (<b>g</b>) Measurement range and error for the three wavelengths of <math display="inline"><semantics> <mrow> <msub> <mi>λ</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>λ</mi> <mn>2</mn> </msub> <mo>,</mo> <msub> <mi>λ</mi> <mn>3</mn> </msub> </mrow> </semantics></math>. Courtesy of Jianji Dong [<a href="#B102-photonics-06-00013" class="html-bibr">102</a>].</p>
Full article ">Figure 13
<p>(<b>a</b>) Schematic of the silicon phase shifted WBG (PS-WBG) employed as a linear-optics frequency discriminator reported in [<a href="#B103-photonics-06-00013" class="html-bibr">103</a>]. TX port, transmission port; RX port, reflection port. Simulated (dashed line) and measured (solid line); (<b>b</b>) linear optical transmission and (<b>c</b>) reflection spectral responses of the PS-WBG; (<b>d</b>,<b>e</b>) zoom with overlapped OSSB+C spectrum. (<b>f</b>) Estimated frequency (red dots) and corresponding error (blue dots).</p>
Full article ">Figure 14
<p>(<b>a</b>) Schematic of the on-chip four wave mixing (FWM)-based instantaneous frequency measurement (IFM) system reported in [<a href="#B104-photonics-06-00013" class="html-bibr">104</a>]. <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>t</mi> </mrow> </semantics></math>: tunable delay element; BPF: optical bandpass filter. (<b>b</b>) DFWM and NDFWM mixing processes between the two channels. (<b>c</b>) IFM RF system response with <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>t</mi> </mrow> </semantics></math> = 8.3 ps. (<b>d</b>) Frequency estimation measurement over a single 40 GHz frequency band (inset, histogram of the frequency measurement error; rms value = 318.9 MHz). Reconfiguration of the IFM system response between (<b>e</b>) a high-bandwidth/error state (low <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>t</mi> </mrow> </semantics></math>) and (<b>f</b>) a low-bandwidth/error state (high <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>t</mi> </mrow> </semantics></math>)). (<b>g</b>) IFM system response with <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>t</mi> </mrow> </semantics></math> = 69.4 ps. (<b>h</b>) Frequency estimation measurement for six separate 7.2 GHz frequency bands (inset, histogram of the frequency measurement error; rms value = 40.2 MHz). Courtesy of David Marpaung [<a href="#B104-photonics-06-00013" class="html-bibr">104</a>].</p>
Full article ">Figure 15
<p>Operation concept of MPFs based on a single micro-resonator (<b>a</b>) All-pass micro-resonator and its power transmission spectrum. Schematic illustration of the schemes for (<b>b</b>) bandpass and (<b>c</b>) notch MPF implementations based on optical-to-RF mapping concept.</p>
Full article ">Figure 16
<p>General frequency-tuning mechanisms for MPFs. (<b>a</b>) Tuning the optical carrier wavelength using a wavelength-tunable continuous wave (CW) laser. (<b>b</b>) Thermally tuning the resonance wavelength of the micro-resonator. Optically tuning the resonance wavelength of the micro-resonator using adjusting (<b>c</b>) the carrier light, and (<b>d</b>) pump light power levels.</p>
Full article ">Figure 17
<p>Implemented bandpass MPFs based on silicon micro-resonators. (<b>a</b>–<b>d</b>) On-chip integrated MPF based on microdisk resonator (MDR) reported in [<a href="#B61-photonics-06-00013" class="html-bibr">61</a>]: (<b>a</b>) perspective view of the proposed MPF, (<b>b</b>) image of the fabricated silicon photonic chip, schematic view of the thermally tunable MDR, cross-sectional view of the MDR, (<b>c</b>) experimental results of MPF and (<b>d</b>) measured frequency responses tuning from 3 to 10 GHz. Courtesy of and adapted from Jianping Yao [<a href="#B61-photonics-06-00013" class="html-bibr">61</a>]. (<b>e</b>–<b>h</b>) Ultracompact bandpass MPF based on a PhC microcavity reported in [<a href="#B105-photonics-06-00013" class="html-bibr">105</a>]: (<b>e</b>) SEM image, and (<b>f</b>) transmission spectrum of the fabricated silicon PhC microcavity, measured MPF responses at (<b>g</b>) different optical carrier wavelengths and (<b>h</b>) different optical carrier power levels. Courtesy of and adapted from Jian Wang [<a href="#B105-photonics-06-00013" class="html-bibr">105</a>]. (<b>i</b>–<b>k</b>) Ultranarrow bandpass MPF reported in [<a href="#B106-photonics-06-00013" class="html-bibr">106</a>]: (<b>i</b>) layout of the used ultra-high-Q MRR, (<b>j</b>) image of the fabricated MRR and (<b>k</b>) experimental results of tunable bandpass MPF. Courtesy of and adapted from Yuan Yu [<a href="#B106-photonics-06-00013" class="html-bibr">106</a>].</p>
Full article ">Figure 18
<p>Implemented all-optical tunable notch MPF proposed in [<a href="#B108-photonics-06-00013" class="html-bibr">108</a>]: (<b>a</b>) SEM image and (<b>b</b>) measured transmission spectrum of the fabricated MRR with the pump light off and on, (<b>c</b>,<b>d</b>) measured microwave responses of the tunable MPF under different pump powers. Courtesy of and adapted from Jian Wang [<a href="#B108-photonics-06-00013" class="html-bibr">108</a>].</p>
Full article ">Figure 19
<p>(<b>a</b>) Schematic illustration of the notch MPF with rejection ratio tunability proposed in [<a href="#B49-photonics-06-00013" class="html-bibr">49</a>], (<b>b</b>) SEM image of the silicon MRR, (<b>c</b>,<b>d</b>) Optical spectra after the tunable optical bandpass filter (OBPF) (dashed lines) and the corresponding MPF responses. Courtesy of and adapted from Jian Wang [<a href="#B49-photonics-06-00013" class="html-bibr">49</a>].</p>
Full article ">Figure 20
<p>(<b>a</b>,<b>b</b>) Bandpass MPF based optical double notch filter reported in [<a href="#B114-photonics-06-00013" class="html-bibr">114</a>]: (<b>a</b>) Schematic diagram of the optical double notch filter, (<b>b</b>) operation principle of the proposed MPF, and (<b>c</b>) measured RF responses. Courtesy of and adapted from Xiaoke Yi [<a href="#B114-photonics-06-00013" class="html-bibr">114</a>]. (<b>d</b>–<b>g</b>) Ultrahigh peak rejection notch MPF proposed in [<a href="#B115-photonics-06-00013" class="html-bibr">115</a>]: (<b>d</b>) The structure and concept of the proposed MPF, (<b>e</b>) optical micrograph of the fabricated device, and (<b>f</b>,<b>g</b>) experimental results. Courtesy of and adapted from Yuan Yu [<a href="#B115-photonics-06-00013" class="html-bibr">115</a>].</p>
Full article ">Figure 21
<p>(<b>a</b>) Working principle of the MPF created by SBS in [<a href="#B118-photonics-06-00013" class="html-bibr">118</a>], (<b>b</b>) setup of the notch filter experiment, (<b>c</b>,<b>d</b>) Measured notch MPF responses. Courtesy of and adapted from Alvaro Casas-Bedoya [<a href="#B118-photonics-06-00013" class="html-bibr">118</a>].</p>
Full article ">Figure 22
<p>Schematic of an AWG system based on spectral shaping and wavelength-to-time mapping (SS-WTT) mapping technique.</p>
Full article ">Figure 23
<p>Recently implemented on-silicon-chip spectral shapers. (<b>a</b>–<b>d</b>) Spectral shaper consisting of five MRRs reported in [<a href="#B22-photonics-06-00013" class="html-bibr">22</a>,<a href="#B140-photonics-06-00013" class="html-bibr">140</a>]: (<b>a</b>) perspective view of the proposed spectral shaper; (<b>b</b>) measured spectral response of the spectral shaper; (<b>c</b>) generated linearly chirped microwave waveforms (LCMW) and (<b>d</b>) its spectrogram. Courtesy of and adapted from Jianping Yao [<a href="#B22-photonics-06-00013" class="html-bibr">22</a>,<a href="#B140-photonics-06-00013" class="html-bibr">140</a>]. (<b>e</b>–<b>i</b>) Spectral shaper incorporating two linearly chirped Bragg gratings (LCBGs) proposed in [<a href="#B22-photonics-06-00013" class="html-bibr">22</a>,<a href="#B137-photonics-06-00013" class="html-bibr">137</a>]: (<b>e</b>) Schematic layout and image of the fabricated on-chip spectral shaper; (<b>f</b>,<b>g</b>) Measured spectral response of the spectral shaper when the offset waveguide length is equal to (<b>f</b>) zero, and (<b>g</b>) the length of the LCBG; (<b>h</b>,<b>i</b>) Experimental results of on-chip spectral shaper corresponding to (<b>f</b>,<b>g</b>), respectively. Courtesy of and adapted from Jianping Yao [<a href="#B22-photonics-06-00013" class="html-bibr">22</a>,<a href="#B137-photonics-06-00013" class="html-bibr">137</a>]. (<b>j</b>–<b>m</b>) Spectral shaper of generation of multiple chirped RF waveforms reported in [<a href="#B138-photonics-06-00013" class="html-bibr">138</a>]: (<b>j</b>,<b>k</b>) proposed spectral shapers based on (<b>j</b>) parallel distributed Fabry-Pérot cavities (DFPCs) and (<b>k</b>) an arrayed waveguide Sagnac interferometer (AWGSI) incorporating LCBGs; (<b>l</b>,<b>m</b>) Experimental results for simultaneous generation of two chirped RF waveforms using the spectral shapers based on (<b>l</b>) DFPCs and (<b>m</b>) AWGSI incorporating LCBGs: top figure shows the measure spectra, middle figure shows the temporal waveforms, and bottom figure shows the calculated spectrograms.</p>
Full article ">Figure 24
<p>Pulse shaper consisting of 8 MRRs reported in [<a href="#B135-photonics-06-00013" class="html-bibr">135</a>]: (<b>a</b>) Operation principle of the proposed pulse shaper; (<b>b</b>) An optical image of a fabricated on-chip pulse shaper; (<b>c</b>) Demonstrated RF bursts and their spectra; (<b>d</b>) Single-tone continuous RF waveforms. Courtesy of and adapted from Minghao Qi and Andrew M. Weiner [<a href="#B135-photonics-06-00013" class="html-bibr">135</a>].</p>
Full article ">Figure 25
<p>(<b>a</b>,<b>b</b>) generation of RF carrier based on a silicon microring modulator reported in [<a href="#B144-photonics-06-00013" class="html-bibr">144</a>]: (<b>a</b>) Schematic diagram of the microring modulator in the RF carrier generation setup; (<b>b</b>) measured electrical spectra for a 13 dBm input microwave signal. Courtesy of and adapted from Hui Yu [<a href="#B144-photonics-06-00013" class="html-bibr">144</a>]. (<b>c</b>–<b>f</b>) proposed silicon optoelectronic oscillators (OEO) reported in [<a href="#B146-photonics-06-00013" class="html-bibr">146</a>]: perspective view of the MDR (<b>c</b>) with a top-placed micro-heater and (<b>d</b>) with a p-type doped micro-heater; experimental result of the OEO with (<b>e</b>) a top-placed micro-heater and (<b>f</b>) a p-type doped silicon heater. Courtesy of and adapted from Jianping Yao [<a href="#B146-photonics-06-00013" class="html-bibr">146</a>].</p>
Full article ">
9 pages, 1637 KiB  
Article
A Terahertz-Microfluidic Chip with a Few Arrays of Asymmetric Meta-Atoms for the Ultra-Trace Sensing of Solutions
by Kazunori Serita, Hironaru Murakami, Iwao Kawayama and Masayoshi Tonouchi
Photonics 2019, 6(1), 12; https://doi.org/10.3390/photonics6010012 - 30 Jan 2019
Cited by 41 | Viewed by 6506
Abstract
Biosensing with terahertz (THz) waves has received large amounts of attention due to its potential to detect the functional expression of biomolecules in a label-free fashion. However, many practical challenges against the diffraction limit of THz waves and the strong absorption of THz [...] Read more.
Biosensing with terahertz (THz) waves has received large amounts of attention due to its potential to detect the functional expression of biomolecules in a label-free fashion. However, many practical challenges against the diffraction limit of THz waves and the strong absorption of THz waves into polar solvents still remain in the development of compact biosensors. Here, we present a non-linear, optical, crystal-based THz-microfluidic chip with a few arrays of asymmetric meta-atoms, an elementary unit of metamaterials, for the measurement of trace amounts of solution samples. A near-field THz emission source, that is locally generated in the process of optical rectification at a fs (femtosecond) laser irradiation spot, induces a sharp Fano resonance and modifies the resonance frequency of the meta-atoms when the channel is filled with solution samples of different concentrations. Using this chip, we successfully detected minute changes in the concentration of trace amounts of mineral water and aqueous sugar solutions by monitoring the shift in the resonance frequency. A higher detectable sensitivity of 1.4 fmol of solute in a 128 pL volume of solution was achieved. This was an improvement of one order of magnitude in the sensitivity compared to our previous experiment. Full article
(This article belongs to the Special Issue Terahertz Photonics)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) A schematic drawing and (<b>b</b>) an optical image of the terahertz (THz) microfluidic chip. (<b>c</b>,<b>d</b>) The simulated THz transmittance spectra of the Fano and quadrupole resonances, respectively.</p>
Full article ">Figure 2
<p>(<b>a</b>) The measured and (<b>b</b>) calculated terahertz (THz) transmittance spectra of the Fano resonance with different values of “<span class="html-italic">d</span>”. (<b>c</b>) The measured and (<b>d</b>) calculated THz transmittance spectra of the quadrupole resonance with different values of “<span class="html-italic">d</span>”.</p>
Full article ">Figure 3
<p>The simulated terahertz (THz) electric field distribution at the resonance frequency when the periods of the meta-atoms are (<b>a</b>,<b>b</b>) 100 μm or (<b>c</b>) 180 μm. The polarization of the THz waves is (<b>a</b>,<b>c</b>) perpendicular or (<b>b</b>) parallel to the gap. The number of meta-atoms is fixed at 5 × 5 units for all cases.</p>
Full article ">Figure 4
<p>The average terahertz (THz) transmittance spectra of (<b>a</b>) commercial mineral water and (<b>b</b>) a glucose water solution at different concentrations. (<b>c</b>) The differential THz transmittance spectra with the deviations from the resonance frequency of the distilled water.</p>
Full article ">
12 pages, 6513 KiB  
Article
A Bio-Compatible Fiber Optic pH Sensor Based on a Thin Core Interferometric Technique
by Magnus Engholm, Krister Hammarling, Henrik Andersson, Mats Sandberg and Hans-Erik Nilsson
Photonics 2019, 6(1), 11; https://doi.org/10.3390/photonics6010011 - 30 Jan 2019
Cited by 11 | Viewed by 4575
Abstract
There is an increasing demand for compact, reliable and versatile sensor concepts for pH-level monitoring within several industrial, chemical as well as bio-medical applications. Many pH sensors concepts have been proposed, however, there is still a need for improved sensor solutions with respect [...] Read more.
There is an increasing demand for compact, reliable and versatile sensor concepts for pH-level monitoring within several industrial, chemical as well as bio-medical applications. Many pH sensors concepts have been proposed, however, there is still a need for improved sensor solutions with respect to reliability, durability and miniaturization but also for multiparameter sensing. Here we present a conceptual verification, which includes theoretical simulations as well as experimental evaluation of a fiber optic pH-sensor based on a bio-compatible pH sensitive material not previously used in this context. The fiber optic sensor is based on a Mach-Zehnder interferometric technique, where the pH sensitive material is coated on a short, typically 20-25 mm thin core fiber spliced between two standard single mode fibers. The working principle of the sensor is simulated by using COMSOL Multiphysics. The simulations are used as a guideline for the construction of the sensors that have been experimentally evaluated in different liquids with pH ranging from 1.95 to 11.89. The results are promising, showing the potential for the development of bio-compatible fiber optic pH sensor with short response time, high sensitivity and broad measurement range. The developed sensor concept can find future use in many medical- or bio-chemical applications as well as in environmental monitoring of large areas. Challenges encountered during the sensor development due to variation in the design parameters are discussed. Full article
(This article belongs to the Special Issue Advanced Optical Materials and Devices)
Show Figures

Figure 1

Figure 1
<p>Principle of a single mode-multi mode-single mode (SMS) thin core Mach-Zehnder interferometer. Light from the core of the left SMF28 fiber is coupled to cladding modes in the short (TCF), which can interact with the surrounding liquid depending on the liquid refractive index. For a broadband light source, a constructive or destructive interference pattern is created at the right (TCF)/SMF28 interface, which is guided in the core of the right SMF28 fiber.</p>
Full article ">Figure 2
<p>A COMSOL multiphysics simulation of an inline interferometer at the wavelength 1550 nm and an external RI 1.33. Inset (<b>a</b>) shows the whole simulated structure, whereas the insets (<b>b</b>,<b>c</b>) show the input and output part of the thin core fiber (TCF) respectively.</p>
Full article ">Figure 3
<p>Electric field along a center line running through the fiber.</p>
Full article ">Figure 4
<p>Electric field in the center of fiber, at wavelength 1550 nm and different surrounding refractive indexes, RI 1.33, 1.37 and 1.41. The inset shows an enlargement of the destructive interference around 540 um where a small shift is observed for increasing RI.</p>
Full article ">Figure 5
<p>Typical transmission spectrum measured on a TCF inline sensor with a length of 24.2 mm held in air (<b>a</b>) and center wavelength for the left minima in (<b>a</b>) as a function of TCF length (<b>b</b>).</p>
Full article ">Figure 6
<p>The wavelength of the transmission minima for a 20.0 mm long TCF sensor submerged in solutions with different RI (<b>a</b>) and the sensitivity in nm per RIU as a function of RI (<b>b</b>).</p>
Full article ">Figure 7
<p>TCF coated with a thin, ∼0.5 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m layer of pH-sensitive coating based on 1.3-BDDA and PIP (<b>a</b>). Large volume expansion of the polymer creating problem with adhesion to the glass fiber when the coated fiber is submerged in a liquid (<b>b</b>).</p>
Full article ">Figure 8
<p>Spectral response from a 1.4-BDDA/PIP polymer coated inline sensor when submerged in liquids with pH-levels ranging from 1.95 to 11.89. The wavelength of the central minima as a function of pH is shown in the inset for pH 6.12 to 11.89.</p>
Full article ">Figure 9
<p>Schematic picture showing the interaction of the evanescent wave with the hydrogel coated layer on the fiber cladding. For thin hydrogel layers, the evanescent wave will be influenced not only by the RI of the hydrogel but also the RI of the surrounding liquid (<b>a</b>). For thicker layers the evanescent wave will only be influenced by the effective index of the hydrogel (<b>b</b>).</p>
Full article ">Figure 10
<p>Wavelength of the central minima as a function of time for a sensor repeatedly submerged in pH-levels of 7.10 and 8.15.</p>
Full article ">
8 pages, 1087 KiB  
Article
Nearly Single-Cycle Terahertz Pulse Generation in Aperiodically Poled Lithium Niobate
by Yuri Avetisyan and Masayoshi Tonouchi
Photonics 2019, 6(1), 9; https://doi.org/10.3390/photonics6010009 - 27 Jan 2019
Cited by 6 | Viewed by 3576
Abstract
In the present work, an opportunity of nearly single-cycle THz pulse generation in aperiodically poled lithium niobate (APPLN) crystal is studied. A radiating antenna model is used to simulate the THz generation from chirped APPLN crystal pumped by a sequence of femtosecond laser [...] Read more.
In the present work, an opportunity of nearly single-cycle THz pulse generation in aperiodically poled lithium niobate (APPLN) crystal is studied. A radiating antenna model is used to simulate the THz generation from chirped APPLN crystal pumped by a sequence of femtosecond laser pulses with chirped delays (m = 1, 2, 3 …) between adjacent pulses. It is shown that by appropriately choosing Δtm, it is possible to obtain temporal overlap of all THz pulses generated from positive (or negative) domains. This results in the formation of a nearly single-cycle THz pulse if the chirp rate of domain length δ in the crystal is sufficiently large. In the opposite case, a few cycle THz pulses are generated with the number of the cycles depending on δ. The closed-form expression for the THz pulse form is obtained. The peak THz electric field strength of 0.3 MV/cm is predicted for APPLN crystal pumped by a sequence of laser pulses with peak intensities of the separate pulse in the sequence of about 20 GW/cm2. By focusing the THz beam and increasing the pump power, the field strength can reach values in the order of few MV/cm. Full article
(This article belongs to the Special Issue Terahertz Photonics)
Show Figures

Figure 1

Figure 1
<p>The schematic illustration of APPLN crystal, where white and dark colors are used for regions with opposite sign of nonlinear coefficient <span class="html-italic">d</span><sub>33</sub>.</p>
Full article ">Figure 2
<p>The temporal forms of THz pulses generated in the APPLN crystal by original laser pulse (red line) and its replicas delayed at Δ<span class="html-italic">t</span><sub>2</sub> (blue curve) and Δ<span class="html-italic">t</span><sub>3</sub> (green curve). The dashed line indicates nearly perfect overlapping of the fields radiated from the first three positive domains, when they are excited by laser pulses with the delays of Δ<span class="html-italic">t</span><sub>1</sub> = 0, Δ<span class="html-italic">t</span><sub>2</sub>, and Δ<span class="html-italic">t</span><sub>3</sub>.</p>
Full article ">Figure 3
<p>(<b>a</b>) Temporal forms of THz pulses generated in APPLN structures having different <span class="html-italic">δ</span> = 1.6 μm (red curve), <span class="html-italic">δ</span> = 0.8 μm (blue curve) and <span class="html-italic">δ</span> = 0.5 (green curve), respectively. The temporal forms are offset on the ordinate axis for clarity. In each case the crystal is pumped by the sequence of the laser pulses having delays Δ<span class="html-italic">τ<sub>m</sub></span> with <span class="html-italic">m</span> = 1, 2, 3,…16. (<b>b</b>) Format of the pump pulses used for cases <span class="html-italic">δ</span> = 1.6 (red line) and <span class="html-italic">δ</span> = 0.8 (blue line), respectively.</p>
Full article ">Figure 4
<p>The spectra of THz generation in chirped APPLN crystal for the cases of <span class="html-italic">δ</span> = 1.6 µm (red curve) and <span class="html-italic">δ</span> = 0.8 µm (blue curve).</p>
Full article ">
10 pages, 3484 KiB  
Article
Imaging of Chemical Reactions Using a Terahertz Chemical Microscope
by Toshihiko Kiwa, Tatsuki Kamiya, Taiga Morimoto, Kentaro Fujiwara, Yuki Maeno, Yuki Akiwa, Masahiro Iida, Taihei Kuroda, Kenji Sakai, Hidetoshi Nose, Masaki Kobayashi and Keiji Tsukada
Photonics 2019, 6(1), 10; https://doi.org/10.3390/photonics6010010 - 27 Jan 2019
Cited by 14 | Viewed by 3908
Abstract
This study develops a terahertz (THz) chemical microscope (TCM) that visualizes the distribution of chemical reaction on a silicon-based sensing chip. This chip, called the sensing plate, was fabricated by depositing Si thin films on a sapphire substrate and thermally oxidizing the [...] Read more.
This study develops a terahertz (THz) chemical microscope (TCM) that visualizes the distribution of chemical reaction on a silicon-based sensing chip. This chip, called the sensing plate, was fabricated by depositing Si thin films on a sapphire substrate and thermally oxidizing the Si film surface. The Si thin film of the sensing plate was irradiated from the substrate side by a femtosecond laser, generating THz pulses that were radiated into free space through the surface field effect of the Si thin film. The surface field responds to chemical reactions on the surface of the sensing plate, changing the amplitude of the THz pulses. This paper first demonstrates the principle and experimental setup of the TCM and performs the imaging and measurement of chemical reactions, including the reactions of bio-related materials. Full article
(This article belongs to the Special Issue Terahertz Photonics)
Show Figures

Figure 1

Figure 1
<p>Schematic of the sensing plate. The sensing plate consists of a Si thin film sandwiched between a SiO<sub>2</sub> thin film and a sapphire substrate. Femtosecond laser pulses penetrate the sensing plate, generating THz pulses in the Si thin film. The amplitude of the THz pulses depends on the chemical reactions on the surface of the sensing plate.</p>
Full article ">Figure 2
<p>Energy band diagram of the sensing plate. The conduction and valence bands of the Si thin film are slightly bent near the Si–SiO<sub>2</sub> boundary, forming a natural depletion field near the edge of the Si film. When a chemical reaction shifts the electric potential on the sensing-plate surface, it simultaneously changes the magnitude of the depletion field.</p>
Full article ">Figure 3
<p>Relation between the electric potential on the sensing-plate surface and the amplitude of the radiated THz pulses. (<b>a</b>) Schematic of the Ti electrode on the sensing plate; (<b>b</b>) amplitude of the THz pulses as a function of the applied voltage. The black line is the best linear fit to the data.</p>
Full article ">Figure 4
<p>Schematic of the optical setup of the TCM.</p>
Full article ">Figure 5
<p>Photograph of the prototype TCM.</p>
Full article ">Figure 6
<p>(<b>a</b>) Cross-sectional schematic and (<b>b</b>) photograph of the microflow channel on the sensing plate. (<b>c</b>) Microflow channel in (<b>b</b>) superimposed with the distribution of THz radiation from the sensing plate. The color scale is normalized by the maximum amplitude in the image.</p>
Full article ">Figure 7
<p>(<b>a</b>) Cross-sectional schematic of the sensing plate with an ion-selective membrane on the surface. The solution containing sodium ions was dropped onto the membrane. (<b>b</b>) THz pulse amplitude versus sodium-ion concentration in the solution during femtosecond laser illumination beneath the ion-selective membrane. The red line is the linear best fit to the data.</p>
Full article ">Figure 8
<p>(<b>a</b>) Cross-sectional schematic of the sensing plate with antibodies on the surface of the sensing plate and (<b>b</b>) photograph of the solution wells filled with the antigens. Four wells were fabricated on the sensing plate.</p>
Full article ">Figure 9
<p>Amplitude changes in the THz pulses before and after the immune reaction. (<b>a</b>) Amplitude map (dashed lines indicate the well edges) and (<b>b</b>) amplitude change versus concentration of mannose dropped into the wells on the sensing plate.</p>
Full article ">
7 pages, 5431 KiB  
Article
Holography Using Curved Metasurfaces
by James Burch and Andrea Di Falco
Photonics 2019, 6(1), 8; https://doi.org/10.3390/photonics6010008 - 26 Jan 2019
Cited by 14 | Viewed by 5332
Abstract
In this work, we demonstrate nonflat metasurface holograms with applications in imaging, sensing, and anticounterfeiting. For these holograms, the image and its symmetry properties, with respect to the polarization of the light, depend on the specific shape of the substrate. Additionally, the sensitivity [...] Read more.
In this work, we demonstrate nonflat metasurface holograms with applications in imaging, sensing, and anticounterfeiting. For these holograms, the image and its symmetry properties, with respect to the polarization of the light, depend on the specific shape of the substrate. Additionally, the sensitivity of the holographic image to the substrate shape can be engineered by distributing the phase information into determined areas of the metasurface. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Light propagation from <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="bold">r</mi> <mi>O</mi> </msub> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mi>O</mi> </msub> <mo>,</mo> <msub> <mi>y</mi> <mi>O</mi> </msub> <mo>,</mo> <msub> <mi>z</mi> <mi>O</mi> </msub> <mo>)</mo> </mrow> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="bold">r</mi> <mi>I</mi> </msub> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mi>I</mi> </msub> <mo>,</mo> <msub> <mi>y</mi> <mi>I</mi> </msub> <mo>,</mo> <msub> <mi>z</mi> <mi>I</mi> </msub> <mo>)</mo> </mrow> </mrow> </semantics></math> using the Rayleigh–Sommerfeld equation. The holographic image originates from the metasurface (MS) and is projected onto the screen.</p>
Full article ">Figure 2
<p>(<b>a</b>) The nanorods comprising the MS and imaged with SEM after the membrane was lifted from the rigid initial substrate; (<b>b</b>) An experimental image displaying an MS conformed to a nonflat surface. The central patterned area appears darker due to the reduced reflectivity with respect to the surrounding region.</p>
Full article ">Figure 3
<p>Experimental holographic images where (<b>a</b>,<b>b</b>) were illuminated with right-handed (RCP) and (<b>c</b>,<b>d</b>) with left-handed circular polarizations (LCP), respectively, with <math display="inline"><semantics> <mrow> <mi>λ</mi> <mtext> </mtext> <mo>=</mo> <mtext> </mtext> <mn>630</mn> </mrow> </semantics></math> nm. (<b>a</b>,<b>c</b>) used a concave substrate with a radius of curvature of 6 mm; (<b>b</b>,<b>d</b>) used a convex substrate with the same radius of curvature. The LCP images are rotated 180 degrees in the holographic image plane compared to the RCP images.</p>
Full article ">Figure 4
<p>Simulated and experimental holographic image results, comparing MS designed for various convex cylinders with differing radii of curvature. In each case, the MS was analyzed as being applied to a convex cylinder with a radius of curvature of 6 mm and <math display="inline"><semantics> <mrow> <mi>λ</mi> <mtext> </mtext> <mo>=</mo> <mtext> </mtext> <mn>630</mn> </mrow> </semantics></math> nm. For the experimental results, the scale bar represents 10 mm. The numbers correspond to the correlation coefficient, where 1 is perfect correlation.</p>
Full article ">Figure 5
<p>The fabrication steps for our MS. (<b>a</b>) Spinning a lift-off layer and a thick SU-8 manipulation layer; (<b>b</b>) evaporation of two layers of gold spaced by an SU-8 layer; (<b>c</b>) spinning, electron beam exposure, and development of an SU-8 resist layer; (<b>d</b>) dry etching to remove unmasked gold and dissolving the Omnicoat layer.</p>
Full article ">
26 pages, 7761 KiB  
Article
THz Mixing with High-TC Hot Electron Bolometers: A Performance Modeling Assessment for Y-Ba-Cu-O Devices
by Romain Ladret, Annick Dégardin, Vishal Jagtap and Alain Kreisler
Photonics 2019, 6(1), 7; https://doi.org/10.3390/photonics6010007 - 25 Jan 2019
Cited by 7 | Viewed by 4235
Abstract
Hot electron bolometers (HEB) made from high-TC superconducting YBa2Cu3O7x (YBCO) oxide nano-constrictions are promising THz mixers, due to their expected wide bandwidth, large mixing gain, and low intrinsic noise. The challenge for YBCO resides, [...] Read more.
Hot electron bolometers (HEB) made from high-TC superconducting YBa2Cu3O7x (YBCO) oxide nano-constrictions are promising THz mixers, due to their expected wide bandwidth, large mixing gain, and low intrinsic noise. The challenge for YBCO resides, however, in the chemical reactivity of the material and the related aging effects. In this paper, we model and simulate the frequency dependent performance of YBCO HEBs operating as THz mixers. We recall first the main hypotheses of our hot spot model taking into account both the RF frequency effects in the YBCO superconducting transition and the nano-constriction impedance at THz frequencies. The predicted performance up to 4 THz is given in terms of double sideband noise temperature TDSB and conversion gain G. At 2.5 THz for instance, TDSB ≅ 1000 K and G ≅ − 6 dB could be achieved at 12.5 μW local oscillator power. We then consider a standoff target detection scheme and examine the feasibility with YBCO devices. For instance, detection at 3 m through cotton cloth in passive imaging mode could be readily achieved in moderate humidity conditions with 10 K resolution. Full article
(This article belongs to the Special Issue Terahertz Photonics)
Show Figures

Figure 1

Figure 1
<p>Superconducting hot electron bolometers (HEB): (<b>a</b>) Illustration of the point bolometer thermal model with two-reservoirs (electrons and phonons of the superconductor); (<b>b</b>) Sketch of a YBCO HEB constriction (of length <span class="html-italic">L</span>, width <span class="html-italic">w</span>, and thickness <span class="html-italic">θ</span>) connected to the arms of a THz planar antenna.</p>
Full article ">Figure 2
<p>Simplified schematic for HEB standoff THz detection arrangement. The target resolved area is deduced from the Airy pattern at the HEB location (Equation (19)).</p>
Full article ">Figure 3
<p>Frequency-dependent YBCO superconducting transition illustrated by the constriction resistance vs. frequency. The DC plot (<span class="html-italic">f</span> = 0) is a fit from experiment [<a href="#B30-photonics-06-00007" class="html-bibr">30</a>]. The gaussian curves describe the <span class="html-italic">T</span><sub>C</sub> distribution as discussed in the text. Fermi-Dirac fits have been used for the device performance simulation as more appropriate to the convergence of numerical solutions. Redrawn after [<a href="#B27-photonics-06-00007" class="html-bibr">27</a>].</p>
Full article ">Figure 4
<p>For devices A and B, electron temperature <span class="html-italic">T</span><sub>e</sub>(<span class="html-italic">x</span>) (solid curves) and phonon temperature <span class="html-italic">T</span><sub>p</sub>(<span class="html-italic">x</span>) (dashed curves) profiles: (<b>a</b>) For device A (<span class="html-italic">P</span><sub>LO</sub> = 5 μW); (<b>b</b>) for device B (<span class="html-italic">P</span><sub>LO</sub> = 35 μW). <span class="html-italic">T</span><sub>0</sub> and <span class="html-italic">T</span><sub>C</sub> are the reference (cold finger) and mid-transition critical temperatures, respectively. Arrows on <span class="html-italic">T</span><sub>C</sub> lines delimit the hot spot regions Δ<span class="html-italic">x</span><sup>HS</sup> (see text). Curve labels indicate the DC bias current <span class="html-italic">I</span><sub>0</sub> values.</p>
Full article ">Figure 5
<p>For devices A and B, DC current vs. DC voltage maps according to <span class="html-italic">P</span><sub>LO</sub> values, at 400 GHz LO frequency: (<b>a</b>) For device A (<span class="html-italic">T</span><sub>0</sub> = 60 Κ); (<b>b</b>) for device B (<span class="html-italic">T</span><sub>0</sub> = 70 Κ). Redrawn after [<a href="#B45-photonics-06-00007" class="html-bibr">45</a>].</p>
Full article ">Figure 6
<p>For the NbN HEB reported in Reference [<a href="#B24-photonics-06-00007" class="html-bibr">24</a>], the comparison between <span class="html-italic">I</span>-<span class="html-italic">V</span> plots measured on the fabricated device and simulation according to the model developed in Reference [<a href="#B24-photonics-06-00007" class="html-bibr">24</a>] and our present model.</p>
Full article ">Figure 7
<p>For an HEB constriction of dimensions <span class="html-italic">L</span> = <span class="html-italic">w</span> = 400 nm and <span class="html-italic">θ</span> = 35 nm (<a href="#photonics-06-00007-f001" class="html-fig">Figure 1</a>), maps exhibiting double sideband noise temperature <span class="html-italic">T</span><sub>DSB</sub> levels in DC bias power vs. LO power coordinates. Impedance matching coefficient <span class="html-italic">α</span><sub>imp</sub> with the antenna was included.</p>
Full article ">Figure 8
<p>For an HEB constriction of dimensions <span class="html-italic">L</span> = <span class="html-italic">w</span> = 400 nm and <span class="html-italic">θ</span> = 35 nm (<a href="#photonics-06-00007-f001" class="html-fig">Figure 1</a>), maps exhibiting Johnson noise and thermal fluctuation noise contributions to <span class="html-italic">T</span><sub>DSB</sub> levels in DC bias power vs. LO power coordinates. Impedance matching coefficient <span class="html-italic">α</span><sub>imp</sub> with the antenna was included.</p>
Full article ">Figure 9
<p>For an HEB constriction of dimensions <span class="html-italic">L</span> = <span class="html-italic">w</span> = 400 nm and <span class="html-italic">θ</span> = 35 nm (<a href="#photonics-06-00007-f001" class="html-fig">Figure 1</a>), maps exhibiting <span class="html-italic">G</span> levels in DC bias power vs. LO power coordinates. Impedance matching coefficient <span class="html-italic">α</span><sub>imp</sub> with the antenna was included.</p>
Full article ">Figure 10
<p>For devices A and B, conversion gain vs. <span class="html-italic">f</span><sub>IF</sub>, at <span class="html-italic">P</span><sub>LO</sub> values minimizing <span class="html-italic">T</span><sub>DSB</sub>. Data for device A were available in the QS regime [<a href="#B25-photonics-06-00007" class="html-bibr">25</a>], whereas data at 2.5 THz were used for device B [<a href="#B45-photonics-06-00007" class="html-bibr">45</a>].</p>
Full article ">Figure 11
<p>Standoff detection double sideband noise temperature requirements: (<b>a</b>) As a function of operating frequency for various atmospheric humidity contents (at fixed Δ<span class="html-italic">T</span> and <span class="html-italic">d</span><sub>T</sub>); (<b>b</b>) as a function of target distance for various target temperature resolutions (at fixed <span class="html-italic">ν</span><sub>0</sub> and <span class="html-italic">RH</span>).</p>
Full article ">Figure 12
<p>Comparing low-voltage behavior of DC <span class="html-italic">I</span>-<span class="html-italic">V</span> plots taking impedance matching factor with antenna into account: (<b>a</b>) Regular hot-spot model with uniform <span class="html-italic">P</span><sub>LO</sub> power dissipation along the constriction; (<b>b</b>) our hot-spot model approach with <span class="html-italic">I</span><sub>LO</sub> as an input parameter, i.e., non-uniform <span class="html-italic">P</span><sub>LO</sub> power dissipation.</p>
Full article ">Figure 13
<p>For a YBCO HEB, <span class="html-italic">I</span>-<span class="html-italic">V</span> map, with <span class="html-italic">P</span><sub>LO</sub> levels, from our simulation with 650 GHz RF current (non-uniform LO power dissipation). Measurement points, without and with LO power applied at the same frequency, are also indicated (courtesy J. Raasch [<a href="#B48-photonics-06-00007" class="html-bibr">48</a>]).</p>
Full article ">Figure 14
<p>Example of the real and imaginary parts of a total THz impedance of a constriction; the temperature was assumed uniform in this case.</p>
Full article ">Figure 15
<p>For an HEB constriction of dimensions <span class="html-italic">L</span> = <span class="html-italic">w</span> = 400 nm and <span class="html-italic">θ</span> = 35 nm (<a href="#photonics-06-00007-f001" class="html-fig">Figure 1</a>), maps exhibiting double sideband noise temperature <span class="html-italic">T</span><sub>DSB</sub> levels in DC bias power vs. LO power coordinates. Impedance matching coefficient with the antenna was not included (<span class="html-italic">α</span><sub>imp</sub> = constant = 1).</p>
Full article ">Figure 16
<p>For HEB heterodyne receivers or mixers, double sideband noise temperature as a function of operating frequency. DC and PC are for diffusion cooled and phonon cooled devices, respectively. Hot spot (HS) model results are those of <a href="#photonics-06-00007-t002" class="html-table">Table 2</a> (optimized) and Reference [<a href="#B27-photonics-06-00007" class="html-bibr">27</a>] (fixed LO power). QL: Quantum limit h<span class="html-italic">ν</span>/(2k<sub>B</sub>). Redrawn and updated after [<a href="#B56-photonics-06-00007" class="html-bibr">56</a>] and [<a href="#B45-photonics-06-00007" class="html-bibr">45</a>].</p>
Full article ">Figure 17
<p>Standoff detection DSB requirements: (<b>a</b>) Required noise temperature as a function of <span class="html-italic">ν</span><sub>0</sub> for various distances at specified Δ<span class="html-italic">T</span> and fixed humidity; simulated <span class="html-italic">T</span><sub>DSB</sub> values for device B are also shown at both optimal <span class="html-italic">P</span><sub>LO</sub> conditions - solid curve (<a href="#photonics-06-00007-t002" class="html-table">Table 2</a>) and at fixed <span class="html-italic">P</span><sub>LO</sub> = 9 µW - dashed curve [<a href="#B27-photonics-06-00007" class="html-bibr">27</a>]; (<b>b</b>) Required distance vs. temperature difference relationship to achieve <span class="html-italic">T</span><sub>DSB</sub> = 1000 K or 2000 K, at fixed <span class="html-italic">RH</span>. Symbols: Computed values, dotted curves: Best fits (according to functions indicated).</p>
Full article ">
13 pages, 6890 KiB  
Article
Multi-Spectral Quantum Cascade Lasers on Silicon With Integrated Multiplexers
by Eric J. Stanton, Alexander Spott, Jon Peters, Michael L. Davenport, Aditya Malik, Nicolas Volet, Junqian Liu, Charles D. Merritt, Igor Vurgaftman, Chul Soo Kim, Jerry R. Meyer and John E. Bowers
Photonics 2019, 6(1), 6; https://doi.org/10.3390/photonics6010006 - 24 Jan 2019
Cited by 11 | Viewed by 5388
Abstract
Multi-spectral midwave-infrared (mid-IR) lasers are demonstrated by directly bonding quantum cascade epitaxial gain layers to silicon-on-insulator (SOI) waveguides with arrayed waveguide grating (AWG) multiplexers. Arrays of distributed feedback (DFB) and distributed Bragg-reflection (DBR) quantum cascade lasers (QCLs) emitting at ∼4.7 µm wavelength are [...] Read more.
Multi-spectral midwave-infrared (mid-IR) lasers are demonstrated by directly bonding quantum cascade epitaxial gain layers to silicon-on-insulator (SOI) waveguides with arrayed waveguide grating (AWG) multiplexers. Arrays of distributed feedback (DFB) and distributed Bragg-reflection (DBR) quantum cascade lasers (QCLs) emitting at ∼4.7 µm wavelength are coupled to AWGs on the same chip. Low-loss spectral beam combining allows for brightness scaling by coupling the light generated by multiple input QCLs into the fundamental mode of a single output waveguide. Promising results are demonstrated and further improvements are in progress. This device can lead to compact and sensitive chemical detection systems using absorption spectroscopy across a broad spectral range in the mid-IR as well as a high-brightness multi-spectral source for power scaling. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Top-view schematic of the QCL array and arrayed waveguide grating (AWG). Mirrors are defined under the red III-V QCL ridges for both the distributed feedback (DFB) and distributed Bragg-reflection (DBR) type lasers. (<b>b</b>) Micrograph of a multi-spectral DFB laser, showing the individual lasers on the left and the AWG combiner on the right.</p>
Full article ">Figure 2
<p>Processing steps: (<b>a</b>) gratings and Si waveguides are defined; (<b>b</b>) the QCL chip is bonded to the Si wafer; (<b>c</b>) the InP substrate is removed; (<b>d</b>) top contact and cladding layers are dry etched; (<b>e</b>) the active region is wet etched; (<b>f</b>) the bottom contact region is defined and gold is deposited; (<b>g</b>) Si<math display="inline"><semantics> <msub> <mrow/> <mn>3</mn> </msub> </semantics></math>N<math display="inline"><semantics> <msub> <mrow/> <mn>4</mn> </msub> </semantics></math> is deposited and the bottom contact layer is etched; (<b>h</b>) vias are etched and the top contact metal is deposited; (<b>i</b>) probe metal is deposited.</p>
Full article ">Figure 3
<p>SEMs of the (<b>a</b>) star coupler of an AWG, (<b>b</b>) transition between the free-propagation region of the star coupler to the arrayed waveguides of an AWG, (<b>c</b>) slanted view of a QCL taper tip before etching the <span class="html-italic">n</span>-QC structure, where the top layer is the SiO<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math> hardmask, and (<b>d</b>) a grating etched in the top of a Si waveguide before bonding the QCL.</p>
Full article ">Figure 4
<p>Experimental setups for (<b>a</b>) the AWG passive transmission, (<b>b</b>) the QCL LIV characteristics, and (<b>c</b>) the QCL spectral measurements. CS represents a current source and PC represents a polarization controller. (<b>d</b>) A photograph of the QCL array coupled to an AWG with the output collected by DET-A, corresponding to the schematic in (<b>b</b>).</p>
Full article ">Figure 5
<p>Light-current-voltage (LIV) dependence on temperature for (<b>a</b>) DFB and (<b>b</b>) DBR devices.</p>
Full article ">Figure 6
<p>Current density at threshold (in orange) and slope efficiency (in blue) extracted as a function of temperature for (<b>a</b>) DFB and (<b>b</b>) DBR lasers.</p>
Full article ">Figure 7
<p>Normalized spectral dependence on the temperature for (<b>a</b>) a DFB laser from channel 2 with a 200 ns drive pulse width and (<b>b</b>) a DBR laser from channel 5 with a 100 ns drive pulse width. Heating during the pulse limits the linewidth of each mode.</p>
Full article ">Figure 8
<p>(<b>a</b>) LIV and (<b>b</b>) spectral characteristics of two DFB lasers, corresponding to emission from the hybrid III-V/Si facet after one taper was polished off.</p>
Full article ">Figure 9
<p>Transmission spectra for a similar AWG, fabricated separately from the QCL array. The solid curves are simulated and the points are measured data.</p>
Full article ">Figure 10
<p>(<b>a</b>) LIV plots of a multi-spectral three-channel DFB while driving laser channels 3, 5, and 6. (<b>b</b>) Spectra of the three-channel DFB laser (Laser #1) and another two-channel DFB laser (Laser #2).</p>
Full article ">Figure 11
<p>(<b>a</b>) LIV plots of a multi-spectral two-channel DBR while driving laser channels 1 and 7. (<b>b</b>) Spectra of the two-channel DBR laser (Laser #3) and another two-channel DFB laser (Laser #4).</p>
Full article ">Figure 12
<p>Power degradation dependence on proximity for multiple pairs of lasers.</p>
Full article ">
2 pages, 153 KiB  
Editorial
Acknowledgement to Reviewers of Photonics in 2018
by Photonics Editorial Office
Photonics 2019, 6(1), 5; https://doi.org/10.3390/photonics6010005 - 24 Jan 2019
Cited by 1 | Viewed by 1866
Abstract
Rigorous peer-review is the corner-stone of high-quality academic publishing [...] Full article
Previous Issue
Next Issue
Back to TopTop