[go: up one dir, main page]

Next Issue
Volume 3, December
Previous Issue
Volume 3, June
 
 

Photonics, Volume 3, Issue 3 (September 2016) – 10 articles

  • 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:
3260 KiB  
Article
Experimental Study of Light Propagation in Apple Tissues Using a Multispectral Imaging System
by Mohamed Lamine Askoura, Fabrice Vaudelle and Jean-Pierre L’Huillier
Photonics 2016, 3(3), 50; https://doi.org/10.3390/photonics3030050 - 18 Sep 2016
Cited by 12 | Viewed by 4677
Abstract
This work aimed at highlighting the role played by the skin in the light propagation through the apple flesh. A multispectral Visible-Near Infrared (Vis-NIR) steady-state imaging setup based on the use of four continuous laser sources (633, 763, 784, and 852 nm) and [...] Read more.
This work aimed at highlighting the role played by the skin in the light propagation through the apple flesh. A multispectral Visible-Near Infrared (Vis-NIR) steady-state imaging setup based on the use of four continuous laser sources (633, 763, 784, and 852 nm) and a charge–coupled–device (CCD) camera was developed to record light diffusion inside apple tissues. Backscattering images and light reflectance profiles were studied to reveal optical features of three whole and half-cut apple varieties with and without skin. The optical absorption and scattering properties (μa, μ’s) of intact apples and peeled apples were also retrieved in reflectance mode, using an optimal sensing range of 2.8–10 mm. A relative difference for Δμa ranging from 3.4% to 24.7% was observed for intact apples with respect to peeled apples. Under the same conditions, no significant changes were noted for Δμ’s, which ranged from 0.1% to 1.7%. These findings show that the apple skin cannot be ignored when using Vis-NIR optical imaging as a non-destructive sensing means to reveal major quality attributes of fruits. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental setup used for recording diffuse reflectance images from (<b>a</b>) a whole apple sample; and (<b>b</b>) a half-cut apple sample.</p>
Full article ">Figure 2
<p>(<b>a</b>) Original image captured from Golden Delicious irradiated at the wavelength of 852 nm; (<b>b</b>) specular light removed; (<b>c</b>) RGB color image; (<b>d</b>) processed image; (<b>e</b>) fits of the average profiles as calculated from image (<b>d</b>) with Equation (2); and (<b>f</b>) typical curve fitting using Equation (3).</p>
Full article ">Figure 3
<p>Processed backscattering images. (<b>a</b>,<b>b</b>) Granny Smith with and without skin at 633 nm; (<b>c</b>,<b>d</b>) Granny Smith with and without skin at 852 nm; (<b>e</b>,<b>f</b>) Golden Delicious with and without skin at 763 nm; and (<b>g</b>,<b>h</b>) Royal Gala with and without skin at 784 nm.</p>
Full article ">Figure 4
<p>Reflectance profiles acquired on apple samples with and without skin. (<b>a</b>) Granny Smith at 633 nm; (<b>b</b>) Golden Delicious at 763 nm; (<b>c</b>) Royal Gala at 784 nm; and (<b>d</b>) Granny Smith at 852 nm.</p>
Full article ">Figure 5
<p>Evolution of the ratios (<b>a</b>) <span class="html-italic">α<sub>R</sub></span> and (<b>b</b>) <span class="html-italic">α<sub>F</sub></span> as a function of the wavelength, for three apple cultivars with and without skin.</p>
Full article ">Figure 6
<p>Backscattering images acquired over the cut equatorial plane of a Granny Smith illuminated by a light source located at <span class="html-italic">z</span> = −3 mm, and emitting at 633 nm. (<b>a</b>) Apple with skin; (<b>b</b>) apple without skin; and (<b>c</b>) extracted reflectance profiles from the backscattering images.</p>
Full article ">Figure 7
<p>Effect of the source location on the reflectance profiles as measured on the equatorial plane of a half-cut Royal Gala, with and without skin and irradiated at 633 nm.</p>
Full article ">Figure 8
<p>Plots of the total reflection ratio <math display="inline"> <semantics> <mrow> <msubsup> <mi>α</mi> <mi>R</mi> <mo>*</mo> </msubsup> </mrow> </semantics> </math> as a function of the wavelength, for three apple cultivars with and without skin.</p>
Full article ">Figure 9
<p>(<b>a</b>) Web diagrams related to the five ratios (<span class="html-italic">α<sub>R</sub></span>, <span class="html-italic">α</span>*<span class="html-italic"><sub>R</sub></span>, <span class="html-italic">α<sub>M</sub></span>, <span class="html-italic">α</span>*<span class="html-italic"><sub>M</sub></span>, <span class="html-italic">α<sub>&lt;R&gt;</sub></span>) for the three apple cultivars Royal Gala, Golden Delicious and Granny Smith tested at four wavelengths 633, 763, 784, and 852 nm; (<b>b</b>) Pigment index (<span class="html-italic">CI</span>) estimated with and without skin.</p>
Full article ">
2223 KiB  
Article
Low Loss Electro-Optic Polymer Based Fast Adaptive Phase Shifters Realized in Silicon Nitride and Oxynitride Waveguide Technology
by Lars Baudzus and Peter M. Krummrich
Photonics 2016, 3(3), 49; https://doi.org/10.3390/photonics3030049 - 26 Aug 2016
Cited by 10 | Viewed by 5946
Abstract
We present a comprehensive study on how to design and fabricate low loss electro-optic phase shifters based on an electro-optic polymer and the silicon nitride and silicon oxynitride waveguide material systems. The loss mechanisms of phase shifters with an electro-optic (EO) polymer cladding [...] Read more.
We present a comprehensive study on how to design and fabricate low loss electro-optic phase shifters based on an electro-optic polymer and the silicon nitride and silicon oxynitride waveguide material systems. The loss mechanisms of phase shifters with an electro-optic (EO) polymer cladding are analyzed in detail and design solutions to achieve lowest losses are presented. In order to verify the low loss design a proof of concept prototype phase shifter was fabricated, which exhibits an attenuation of 0.8 dB/cm at 1550 nm and an electro-optic efficiency factor of 27%. Furthermore, the potential of this class of phase shifters is evaluated in numerical simulations, from which the optimal design parameters and achievable figures of merit were derived. The presented phase shifter design has its potential for application in fast adaptive multi stage devices for optical signal processing. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Straightforward design of an electro-optic (EO) polymer based phase shifter, where the EO polymer is used as cladding material.</p>
Full article ">Figure 2
<p>Optimized waveguide design, which enables a more homogenous poling field to reduce poling induced losses.</p>
Full article ">Figure 3
<p>Refractive index matched SiON cladding, which reduces transition losses between poled and non-poled areas and can also be used as bridges over the waveguides in order to connect the electrodes with the contact pads.</p>
Full article ">Figure 4
<p>SEM picture of the active region of a fabricated prototype device.</p>
Full article ">Figure 5
<p>Diagram of the set-up for the study of the EO activity.</p>
Full article ">Figure 6
<p>SEM picture, which illustrates the low sidewall roughness resulting from an optimized etching process.</p>
Full article ">Figure 7
<p>Results of the simulations to determine the optimal core refractive index.</p>
Full article ">Figure 8
<p>Results of the simulations to optimize the core dimensions at a core refractive index of 1.99 and an EO polymer refractive index of 1.7.</p>
Full article ">Figure 9
<p>Simulation results, which show <math display="inline"> <semantics> <mrow> <msub> <mi>V</mi> <mi mathvariant="sans-serif">π</mi> </msub> <mo>·</mo> <mi>L</mi> <mo>·</mo> <msub> <mi>α</mi> <mtext>dB</mtext> </msub> <mo>·</mo> <msub> <mi>r</mi> <mn>33</mn> </msub> </mrow> </semantics> </math> and the optimal electrode distance in dependence of the EO polymer attenuation for the optimized waveguide dimensions.</p>
Full article ">
1338 KiB  
Communication
Photodynamic Therapy-Induced Microvascular Changes in a Nonmelanoma Skin Cancer Model Assessed by Photoacoustic Microscopy and Diffuse Correlation Spectroscopy
by Daniel J. Rohrbach, Hakeem Salem, Mehmet Aksahin and Ulas Sunar
Photonics 2016, 3(3), 48; https://doi.org/10.3390/photonics3030048 - 5 Aug 2016
Cited by 13 | Viewed by 5056
Abstract
One of the main mechanisms of action for photodynamic therapy (PDT) is the destruction of tumor vasculature. We observed the PDT-induced vasculature destruction in a mouse model of skin cancer using two techniques: Photoacoustic microscopy (PAM) and diffuse correlation spectroscopy (DCS). PAM showed [...] Read more.
One of the main mechanisms of action for photodynamic therapy (PDT) is the destruction of tumor vasculature. We observed the PDT-induced vasculature destruction in a mouse model of skin cancer using two techniques: Photoacoustic microscopy (PAM) and diffuse correlation spectroscopy (DCS). PAM showed high-resolution images of the abnormal microvasculature near the establishing tumor area at pre-PDT, as well as the subsequent destruction of those vessels post-PDT. DCS indicated a significant blood flow decrease after PDT, confirming the vascular destruction. Noninvasive assessment of vascular changes may be indicative of therapy response. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Instrument setup. (<b>a</b>) a picture of the PAM setup; (<b>b</b>) a schematic of the setup and; (<b>c</b>) a picture of PDT being performed; (<b>d</b>) shows a picture of the mouse ear with the red area indicating the treatment region and the black dashed line outlining the tumor.</p>
Full article ">Figure 2
<p>Results of the resolution test. (<b>a</b>) shows group 5, elements 1–6 of the 1951 USAF target; (<b>b</b>) shows the PAM image of elements 3–6 and (<b>c</b>) shows the PA signal along a representative line of the image.</p>
Full article ">Figure 3
<p>In vivo results. (<b>a</b>) Pre-PDT; (<b>b</b>) one minute and; (<b>c</b>) Post-PDT PAM images with the white bar indicating 500 µm; (<b>d</b>) Vascular area; (<b>e</b>) vessel diameter and; (<b>f</b>) blood flow all show a decrease with treatment time; (<b>g</b>) shows the relationship between vessel diameter and rBF. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
3508 KiB  
Article
Frequency Tuning and Modulation of a Quantum Cascade Laser with an Integrated Resistive Heater
by Kutan Gürel, Stéphane Schilt, Alfredo Bismuto, Yves Bidaux, Camille Tardy, Stéphane Blaser, Tobias Gresch and Thomas Südmeyer
Photonics 2016, 3(3), 47; https://doi.org/10.3390/photonics3030047 - 30 Jul 2016
Cited by 9 | Viewed by 5169
Abstract
We present a detailed experimental investigation of the use of a novel actuator for frequency tuning and modulation in a quantum cascade laser (QCL) based on a resistive integrated heater (IH) placed close to the active region. This new actuator is attractive for [...] Read more.
We present a detailed experimental investigation of the use of a novel actuator for frequency tuning and modulation in a quantum cascade laser (QCL) based on a resistive integrated heater (IH) placed close to the active region. This new actuator is attractive for molecular spectroscopy applications as it enables fast tuning of the QCL wavelength with a minor influence on the optical output power, and is electrically-controlled. Using a spectroscopic setup comprising a low-pressure gas cell, we measured the tuning and modulation properties of a QCL emitting at 7.8 μm as a function of the active region and IH currents. We show that a current step applied to the IH enables the laser frequency to be switched by 500 MHz in a few milliseconds, as fast as for a step of the current in the active region, and limited by heat dissipation towards the laser sub-mount. The QCL optical frequency can be modulated up to ~100 kHz with the IH current, which is one order of magnitude slower than for the QCL current, but sufficient for many spectroscopic applications. We discuss the experimental results using a thermal model of the heat transfer in terms of cascaded low-pass filters and extract the respective cut-off frequencies. Finally, we present a proof-of-principle experiment of wavelength modulation spectroscopy of a N2O transition performed with a modulation of the IH current and show some potential benefits in comparison to QCL current modulation, which results from the reduced associated amplitude modulation. Full article
(This article belongs to the Special Issue Quantum Cascade Lasers - Advances and New Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Electrical connection scheme of the QCL with IH with their respective current source; (<b>b</b>) current-voltage response of the IH at a sub-mount temperature of 25 °C. The data points (light blue) are well approximated by the pure ohmic response described by the dashed dark blue line.</p>
Full article ">Figure 2
<p>Example of N<sub>2</sub>O absorption line used as a frequency discriminator to measure the FM response of the QCL. The P11e line of the vibrational band ν<sub>1</sub> of N<sub>2</sub>O located at 1275.5 cm<sup>−1</sup> was used in this case and is shown here. The absorption spectrum was measured by tuning the QCL or IH current and recording the voltage of the photodiode at the output of the reference gas cell. The current axis was converted into a relative frequency using the separately measured tuning coefficient. The operating point is shown by the red circle and the linear range by the red line with a slope <span class="html-italic">D</span>.</p>
Full article ">Figure 3
<p>(<b>a</b>) N<sub>2</sub>O absorption spectrum measured by tuning the QCL with the electrical power dissipated in the IH (bottom spectrum obtained at <span class="html-italic">T</span> = 25 °C and <span class="html-italic">I</span><sub>QCL</sub> = 450 mA) and corresponding N<sub>2</sub>O absorption spectrum calculated from HITRAN database [<a href="#B13-photonics-03-00047" class="html-bibr">13</a>] for a N<sub>2</sub>O pressure of 10 mbar diluted in air (total pressure of 70 mbar) over a pathlength of 10 cm (upper spectrum). The reference spectrum is inverted for the clarity of the figure. The total pressure in the reference cell that resulted from air contamination arising from the imperfect cell tightness was estimated to be ~70 mbar from comparisons between measured and calculated spectra; (<b>b</b>) corresponding static tuning curve as a function of the IH current (green markers: experimental points; red dashed line: quadratic fit) and associated variation of the optical power (blue markers: Experimental points; blue dashed line: 3rd order polynomial fit). The experimental data of the optical power were extracted from the measured transmission through the N<sub>2</sub>O cell; the gaps in the data result from the presence of N<sub>2</sub>O absorption lines that have been removed.</p>
Full article ">Figure 4
<p>Static tuning coefficients of the QCL measured for temperature, QCL-current and IH-current tuning, respectively (markers: Experimental points; lines: Linear fits).</p>
Full article ">Figure 5
<p>(<b>a</b>) Experimental principle of the QCL frequency step response measurement using an absorption line of N<sub>2</sub>O; (<b>b</b>) the side of the absorption profile (blue curve) corresponding to the P11e line of N<sub>2</sub>O at 1275.5 cm<sup>−1</sup> acts as a frequency discriminator (dashed red line) that linearly converts the change of the laser frequency into a change of the transmitted optical power detected by a photodiode. A current step of the QCL, IH or TEC was applied (from <span class="html-italic">I</span><sub>1</sub> to <span class="html-italic">I</span><sub>2</sub>, see inset) to produce an exponential frequency change (schematized by the green line) of approximately 500 MHz that was recorded by an oscilloscope.</p>
Full article ">Figure 6
<p>Temporal evolution of the QCL optical frequency measured for a step change of the TEC temperature of ~0.16 K, displayed in a semi-log scale. The inset shows a linear representation. The experimental data are well approximated by a single exponential decay with a time constant of ~100 s (dashed-dotted line).</p>
Full article ">Figure 7
<p>Temporal evolution of the QCL optical frequency measured for a step change of the QCL current (<b>a</b>) and IH current (<b>b</b>), displayed in a semi-log scale. The insets show a linear representation of the initial 0.5-ms decay. The temporal response does not correspond to a single exponential decay of time constant τ<sub>2</sub> (dashed-dotted lines), but is modelled by the sum of four decaying exponential terms (dashed lines).</p>
Full article ">Figure 8
<p>FM transfer function in amplitude (<b>a</b>) and phase (<b>b</b>) obtained for a modulation of the QCL-current (blue lines) and IH-current (red lines), both at different IH currents I<sub>IH</sub>.</p>
Full article ">Figure 9
<p>Relative FM response of the QCL obtained under direct modulation of the injection current (blue points) or IH current (red points) and fitted model (solid lines). Both amplitude (<b>a</b>) and phase (<b>b</b>) responses are shown. A global phase shift of 180° was added to all phase plots.</p>
Full article ">Figure 10
<p>(<b>a</b>) Comparison of the first harmonic WMS signal of a N<sub>2</sub>O line obtained for QCL injection current and IH current modulation (at <span class="html-italic">T</span> ≈ 20 °C and <span class="html-italic">I</span><sub>QCL</sub> ≈ 420 mA), showing the strongly reduced background offset obtained for IH current modulation; (<b>b</b>) ratio between the offset and the amplitude of the 1<span class="html-italic">f</span> signal obtained for QCL and IH modulation at different QCL (IH) currents, converted into a corresponding frequency detuning. For each value of the QCL (IH) current, the laser temperature was slightly varied to keep the laser frequency around the considered N<sub>2</sub>O transition. At each DC current, the amplitude of the applied modulation was adjusted to maximize the peak-to-peak amplitude of the 1<span class="html-italic">f</span> signal.</p>
Full article ">
1680 KiB  
Article
Ar+-Implanted Si-Waveguide Photodiodes for Mid-Infrared Detection
by Brian Souhan, Christine P. Chen, Ming Lu, Aaron Stein, Hassaram Bakhru, Richard R. Grote, Keren Bergman, William M. J. Green and Richard M. Osgood
Photonics 2016, 3(3), 46; https://doi.org/10.3390/photonics3030046 - 27 Jul 2016
Cited by 4 | Viewed by 4223
Abstract
Complementary metal-oxide-semiconductor (CMOS)-compatible Ar+-implanted Si-waveguide p-i-n photodetectors operating in the mid-infrared (2.2 to 2.3 µm wavelengths) are demonstrated at room temperature. Responsivities exceeding 21 mA/W are measured at a 5 V reverse bias with an estimated internal quantum efficiency of 3.1%–3.7%. [...] Read more.
Complementary metal-oxide-semiconductor (CMOS)-compatible Ar+-implanted Si-waveguide p-i-n photodetectors operating in the mid-infrared (2.2 to 2.3 µm wavelengths) are demonstrated at room temperature. Responsivities exceeding 21 mA/W are measured at a 5 V reverse bias with an estimated internal quantum efficiency of 3.1%–3.7%. The dark current is found to vary from a few nanoamps down to less than 11 pA after post-implantation annealing at 350 °C. Linearity is demonstrated over four orders of magnitude, confirming a single-photon absorption process. The devices demonstrate a higher thermal processing budget than similar Si+-implanted devices and achieve higher responsivity after annealing up to 350 °C. Full article
(This article belongs to the Special Issue Advanced Photodetectors Devices and Technologies)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Cross-sectional diagram of the waveguide photodetector with a mode intensity profile for a wavelength of 2.2 µm; (<b>b</b>) Cross section at center of waveguide showing the overlap of the normalized mode intensity with both the defect density and the ion density as calculated using Stopping Range of Ions in Matter (SRIM) (and prior to any annealing or post processing); (<b>c</b>) Waveguide top view indicating the Si waveguide, the Ar<sup>+</sup>-implant region, doped “wings,” and at the end, the fan-out tapered coupler.</p>
Full article ">Figure 2
<p>(<b>a</b>) Photocurrent and dark current for devices annealed at 300 °C and 350 °C; (<b>b</b>) Linearity for devices annealed at 250 °C and 350 °C.</p>
Full article ">Figure 3
<p>(<b>a</b>) Responsivity as a function of bias and annealing temperature for a 250 µm Ar<sup>+</sup>-implanted PD. Significant increases in responsivity were observed after annealing at 150 °C, 200 °C, and 300 °C; (<b>b</b>) Responsivity at a 5 V reverse bias as a function of annealing temperature for various Ar<sup>+</sup>-implanted photodetectors (PDs) and Si<sup>+</sup>-implanted PDs. Unlike the peak seen in responsivity for Si<sup>+</sup>-implanted devices, the responsivity of Ar<sup>+</sup> devices increases with higher temperature annealing [<a href="#B12-photonics-03-00046" class="html-bibr">12</a>]. Due to Al reflow as discussed in the text, we are unable to determine the maximum annealing temperature.</p>
Full article ">Figure 4
<p>Measured Responsivity for the 250 µm PD versus wavelength after high-temperature annealing.</p>
Full article ">
13829 KiB  
Review
Applying Quantum Cascade Laser Spectroscopy in Plasma Diagnostics
by Jürgen Röpcke, Paul B. Davies, Stephan Hamann, Mario Hannemann, Norbert Lang and Jean-Pierre H. Van Helden
Photonics 2016, 3(3), 45; https://doi.org/10.3390/photonics3030045 - 25 Jul 2016
Cited by 23 | Viewed by 7010
Abstract
The considerably higher power and wider frequency coverage available from quantum cascade lasers (QCLs) in comparison to lead salt diode lasers has led to substantial advances when QCLs are used in pure and applied infrared spectroscopy. Furthermore, they can be used in both [...] Read more.
The considerably higher power and wider frequency coverage available from quantum cascade lasers (QCLs) in comparison to lead salt diode lasers has led to substantial advances when QCLs are used in pure and applied infrared spectroscopy. Furthermore, they can be used in both pulsed and continuous wave (cw) operation, opening up new possibilities in quantitative time resolved applications in plasmas both in the laboratory and in industry as shown in this article. However, in order to determine absolute concentrations accurately using pulsed QCLs, careful attention has to be paid to features like power saturation phenomena. Hence, we begin with a discussion of the non-linear effects which must be considered when using short or long pulse mode operation. More recently, cw QCLs have been introduced which have the advantage of higher power, better spectral resolution and lower fluctuations in light intensity compared to pulsed devices. They have proved particularly useful in sensing applications in plasmas when very low concentrations have to be monitored. Finally, the use of cw external cavity QCLs (EC-QCLs) for multi species detection is described, using a diagnostics study of a methane/nitrogen plasma as an example. The wide frequency coverage of this type of QCL laser, which is significantly broader than from a distributed feedback QCL (DFB-QCL), is a substantial advantage for multi species detection. Therefore, cw EC-QCLs are state of the art devices and have enormous potential for future plasma diagnostic studies. Full article
(This article belongs to the Special Issue Quantum Cascade Lasers - Advances and New Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Principal sketch of the optical set-up used for temperature determination. The radiation of the QCLAS system was passed three times through the Pyrex discharge cell. The cell was 60 cm in length with an inner diameter of 20 mm. The experiments were performed under static gas conditions, <span class="html-italic">P<sub>initial</sub></span> = 1.33 mbar [<a href="#B50-photonics-03-00045" class="html-bibr">50</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 2
<p>Detected NO spectrum in the time domain measured by the QCLAS system. The spectrum (solid line) corresponds to 0.8<span class="html-italic">%</span> NO in air at a pressure of 1.33 mbar with an absorption length <span class="html-italic">L</span> = 180 cm. The dashed curve shows the baseline. The embedded diagram shows a simulated NO spectrum using the HITRAN database [<a href="#B50-photonics-03-00045" class="html-bibr">50</a>,<a href="#B83-photonics-03-00045" class="html-bibr">83</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 3
<p>Calculated temporal evolution of the temperature of NO diluted in air (symbols) for a 150 mA plasma pulse. Exponential functions have been fitted to the temperature rise during the pulse, <span class="html-italic">t</span><sub>2</sub>, as well as to the cooling period after the plasma pulse, <span class="html-italic">t</span><sub>3</sub> [<a href="#B50-photonics-03-00045" class="html-bibr">50</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 4
<p>Calculated gas temperatures during the plasma pulse as a function of the mean plasma current. The total pressure of the gas mixture containing 0.8<span class="html-italic">%</span> NO in air was 1.33 mbar [<a href="#B50-photonics-03-00045" class="html-bibr">50</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 5
<p>Schematic of the Pyrex tube coated internally with a TiO<sub>2</sub> sol-gel film and TiO<sub>2</sub> nano-particles, with two electrodes, placed outside the tube [<a href="#B51-photonics-03-00045" class="html-bibr">51</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 6
<p>Time evolution of the concentrations of C<sub>2</sub>H<sub>2</sub> (upper panel) and of CO<sub>2</sub> (lower panel) in phases 3 and 4 (initial gas mixture: 1% C<sub>2</sub>H<sub>2</sub> in Ar, <span class="html-italic">p</span> = 2.6 mbar). After-treatment: ∆—heating to 350 °C and ■—UV radiation exposure [<a href="#B51-photonics-03-00045" class="html-bibr">51</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 7
<p>Photo-catalytic decomposition (i) of CO (○) and formation of CO<sub>2</sub> (●) in phases 3 and 4 after an Ar plasma pre-treatment in phase 1 and (ii) decomposition of CO (<b>□</b>) and formation of CO<sub>2</sub> (■) after an O<sub>2</sub> plasma pre-treatment in phase 1 (Ar RF plasma: <span class="html-italic">p</span> = 0.26 mbar, <math display="inline"> <semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics> </math> = 14 sccm, <span class="html-italic">t</span> = 40 min, O<sub>2</sub> RF plasma: <span class="html-italic">p</span> = 0.53 mbar, <math display="inline"> <semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics> </math> = 14 sccm, <span class="html-italic">t</span> = 30 min, initial gas mixture: 1% CO in Ar, <span class="html-italic">p</span> = 1.3 mbar, UV light exposure in phase 4) [<a href="#B51-photonics-03-00045" class="html-bibr">51</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 8
<p>Experimental set-up used to study surface vibrational relaxation of N<sub>2</sub> [<a href="#B52-photonics-03-00045" class="html-bibr">52</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 9
<p>Time evolution of the absorption signals of CO<sub>2</sub> and N<sub>2</sub>O measured simultaneously in a pulsed dc discharge at <span class="html-italic">p</span> = 133 Pa, <span class="html-italic">I</span> = 50 mA, <span class="html-italic">τ</span> = 5 ms. A fit of the experimental data is also shown. [<a href="#B52-photonics-03-00045" class="html-bibr">52</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 10
<p>Characteristic relaxation frequencies (1/<span class="html-italic">τ</span><sub>rel</sub>) as a function of the concentration of IR tracers left after the discharge pulse. The solid lines show a linear fit of the experimental data [<a href="#B52-photonics-03-00045" class="html-bibr">52</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 11
<p>Effective relaxation frequency 1/<span class="html-italic">τ</span><sub>rel</sub> in a silica reactor, pretreated with a N<sub>2</sub>, O<sub>2</sub> or argon plasma, as a function of the concentration of the IR tracer N<sub>2</sub>O [<a href="#B52-photonics-03-00045" class="html-bibr">52</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 12
<p>Experimental set-up. OAP: off-axis parabolic mirror; QCL: quantum cascade laser; <span class="html-italic">R</span><sub>meas</sub> = 18.4 Ω; <span class="html-italic">U</span><sub>HV</sub>: HV pulse generator; <span class="html-italic">U</span><sub>meas</sub>: HV probe for voltage measurement. The electrodes are represented by the two black rods inside the T-shaped holder [<a href="#B64-photonics-03-00045" class="html-bibr">64</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 13
<p>(<b>a</b>) Temporal evolution of selected atomic species and (<b>b</b>) of electronically excited nitrogen molecules for the case of a current of 150 mA [<a href="#B64-photonics-03-00045" class="html-bibr">64</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 14
<p>Temporal evolution of the measured (symbols) and calculated (full line) NO density in synthetic air for <span class="html-italic">I</span> = 150 mA. The dashed line represents the gas temperature. The dashed-dotted line is the plasma pulse [<a href="#B64-photonics-03-00045" class="html-bibr">64</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 15
<p>Experimental arrangement of the inductively coupled plasma (ICP) reactor, Oxford Plasmalab system 100, combined with a Q-MACS Process fiber system containing two pulsed QCLs [<a href="#B58-photonics-03-00045" class="html-bibr">58</a>], © Springer Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 16
<p>Time dependent CO concentration with and without O<sub>2</sub> plasma pretreatment (CF<sub>4</sub> plasma, <math display="inline"> <semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics> </math> = 12 sccm, <span class="html-italic">p</span> = 1.6 Pa). (Pretreatment: <span class="html-italic">t</span> = 3 min, <span class="html-italic">p</span> = 1.3 Pa, <span class="html-italic">P</span> = 1.5 kW, <math display="inline"> <semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics> </math><sub>O2</sub> = 30 sccm) [<a href="#B58-photonics-03-00045" class="html-bibr">58</a>], © Springer Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 17
<p>Production rate of SiF<sub>4</sub> and the etching rate of SiCOH dependence on power (CF<sub>4</sub> plasma, <math display="inline"> <semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics> </math> = 12 sccm, <span class="html-italic">p</span> = 0.93 Pa) [<a href="#B58-photonics-03-00045" class="html-bibr">58</a>], © Springer Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 18
<p>Schematics of a microscale atmospheric pressure RF plasma jet (<span class="html-italic">μ</span>-APPJ) source [<a href="#B65-photonics-03-00045" class="html-bibr">65</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 19
<p>Set-up of the <span class="html-italic">μ</span>-APPJ with adjacent absorption cell [<a href="#B65-photonics-03-00045" class="html-bibr">65</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 20
<p>Absolute concentration of NO <b>(blue</b> square) and N<sub>2</sub>O (<b>red</b> circle) produced by the micro plasma jet as a function of the absorbed power for a helium flow of 1.4 slm with constant 1400 ppm O<sub>2</sub> and 7100 ppm N<sub>2</sub> admixture. The dashed blue line and the dashed dot red line are the polynomial fit of NO and N<sub>2</sub>O densities, respectively [<a href="#B65-photonics-03-00045" class="html-bibr">65</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 21
<p>Comparison of absolute densities of nitrous species from the <span class="html-italic">μ</span>-APPJ obtained by molecular beam mass spectrometry (MBMS) (for <span class="html-italic">x</span> = 1 and 10 mm) and IR absorption (&gt;100 mm) for a ratio of He/N<sub>2</sub>/O<sub>2</sub> = 99.58/0.35/0.07 at 1.4 slm gas flow. The connecting lines are to show the reader the different decay constants of the species [<a href="#B65-photonics-03-00045" class="html-bibr">65</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 22
<p>Schematic diagram of the spectroscopic experimental set-up based on a QCL absorption diagnostic system implementing the plasma jet. [<a href="#B54-photonics-03-00045" class="html-bibr">54</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 23
<p>Transmittance spectrum of NO<sub>2</sub> produced by the jet at atmospheric pressure combined with a real-time fit [<a href="#B54-photonics-03-00045" class="html-bibr">54</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 24
<p>Absolute concentration of NO<sub>2</sub> produced by the kinpen for different dry air admixtures [<a href="#B54-photonics-03-00045" class="html-bibr">54</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 25
<p>Experimental arrangement of the tunable diode lasers (TDLs), using a cw QCL, and the plasma-enhanced chemical vapor deposition (PECVD) system. Silane is monitored in the reactor volume (D<sub>1</sub>), in the chamber volume (D<sub>2</sub>), and in the pumping line (D<sub>3</sub>) of the PECVD system [<a href="#B57-photonics-03-00045" class="html-bibr">57</a>], Reproduced from [<a href="#B57-photonics-03-00045" class="html-bibr">57</a>], with the permission of AIP Publishing.</p>
Full article ">Figure 26
<p>Plasma termination in a PECVD reactor for process A (<b>a</b>) and process B (<b>b</b>). In process A, Q<sub>SiH4</sub>; Q<sub>H2</sub>, and p were 45 sccm, 1980 sccm, and 4 mbars, respectively. In process B, a pressure of 1 mbar was maintained in the reactor with a pure input SiH<sub>4</sub> flow rate of 120 sccm. Upon termination of the plasma, the SiH<sub>4</sub> recovers its undepleted value with a time constant of 390 ms (<b>a</b>) and 1.82 s (<b>b</b>) (as obtained from an exponential fitting procedure). The time constant is similar to the gas residence time of the respective process [<a href="#B57-photonics-03-00045" class="html-bibr">57</a>], Reproduced from [<a href="#B57-photonics-03-00045" class="html-bibr">57</a>], with the permission of AIP Publishing.</p>
Full article ">Figure 27
<p>Outline of the Oxford Plasmalab System 100 ICP etch chamber equipped with Q-MACS Process fiber system. CF<sub>2</sub> radicals were detected using a multipass cell just above the wafer [<a href="#B59-photonics-03-00045" class="html-bibr">59</a>], Reproduced from [<a href="#B59-photonics-03-00045" class="html-bibr">59</a>], with the permission of AIP Publishing.</p>
Full article ">Figure 28
<p>Time dependent CF<sub>2</sub> concentration while etching blank porous SiCOH (squares) and structured porous SiCOH (circles) using a CF<sub>4</sub> plasma with a rf power of 1000 W, bias power of 60 W, 1.33 Pa total pressure, and a gas flow rate of 25 sccm. The inset shows a schematic diagram of the lateral cross-section of a structured porous SiCOH wafer. The numbers in the upper curve represent the numbers given as the inset [<a href="#B59-photonics-03-00045" class="html-bibr">59</a>], Reproduced from [<a href="#B59-photonics-03-00045" class="html-bibr">59</a>], with the permission of AIP Publishing.</p>
Full article ">Figure 29
<p>Experimental arrangement of the PLANIMOR model reactor combined with QCLAS and OES spectrometers. The plane of the White cell is parallel to the screen and the sample holder (it is rotated by 90° for better illustration) [<a href="#B63-photonics-03-00045" class="html-bibr">63</a>], Reproduced from [<a href="#B63-photonics-03-00045" class="html-bibr">63</a>], with the permission of AIP Publishing.</p>
Full article ">Figure 30
<p>Example of an absorption spectrum at 1356.6 cm<sup>−1</sup> with absorption features of CH<sub>4</sub>, HCN, C<sub>2</sub>H<sub>2</sub> and NH<sub>3</sub> (<span class="html-italic">p</span> = 3 mbar, 10 sccm H<sub>2</sub> + 10 sccm N<sub>2</sub> + 0.5 sccm CH<sub>4</sub>, <span class="html-italic">P</span><sub>screen</sub> = 90 W) [<a href="#B63-photonics-03-00045" class="html-bibr">63</a>], Reproduced from [<a href="#B63-photonics-03-00045" class="html-bibr">63</a>], with the permission of AIP Publishing.</p>
Full article ">Figure 31
<p>Comparing treated (left) and untreated (right) C15 steel samples at <span class="html-italic">t</span> = 4 h, <span class="html-italic">P</span><sub>screen</sub> = 90 W, <math display="inline"> <semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics> </math> = 20 sccm (H<sub>2</sub>:N<sub>2</sub> = 1:1), <span class="html-italic">p</span> = 3 mbar, and <span class="html-italic">T</span><sub>samples</sub> = 823 K [<a href="#B63-photonics-03-00045" class="html-bibr">63</a>], Reproduced from [<a href="#B63-photonics-03-00045" class="html-bibr">63</a>], with the permission of AIP Publishing.</p>
Full article ">Figure 32
<p>The concentrations of NH<sub>3</sub>, CH<sub>4</sub>, HCN and C<sub>2</sub>H<sub>2</sub> depending on the power of the screen plasma (<span class="html-italic">p</span> = 3 mbar, 10 sccm H<sub>2</sub> + 10 sccm N<sub>2</sub> + 0.2 sccm CH<sub>4</sub>) [<a href="#B63-photonics-03-00045" class="html-bibr">63</a>], Reproduced from [<a href="#B63-photonics-03-00045" class="html-bibr">63</a>], with the permission of AIP Publishing.</p>
Full article ">Figure 33
<p>Experimental arrangement of the distributed antenna array (DAA) microwave reactor combined with the AS spectrometer consisting of the IRMA system and an external cavity QCL (EC-QCL) and the OES unit [<a href="#B66-photonics-03-00045" class="html-bibr">66</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 34
<p>CO band stick spectrum containing 31 ground and hot band lines measured with an EC-QCL spectrometer. The intensities of the lines are normalized. In general, the intensities of the hot band lines are about 20 times smaller than the ground state lines, <span class="html-italic">p</span> = 0.35 mbar, <span class="html-italic">P</span> = 3 kW [<a href="#B66-photonics-03-00045" class="html-bibr">66</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 35
<p>Experimental data of absorption lines of CO, R(1) of the ground state and R(17) of the first hot band, together with Gaussian fits for temperature determination. The lines are superposed in position and normalized in intensity in order to compare their relative broadening, <span class="html-italic">p</span> = 0.35 mbar, <span class="html-italic">P</span> = 3 kW, <math display="inline"> <semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics> </math> = 50 sccm, CH<sub>4</sub> admixture: 2.5%, CO<sub>2</sub> admixture: 1% [<a href="#B66-photonics-03-00045" class="html-bibr">66</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 36
<p>Gas temperature T<sub>gas</sub> determined from the Doppler broadening of CO lines as a function of their rotational quantum number for the R and P branches for the ground and the first, second and third excited vibrational levels, <span class="html-italic">p</span> = 0.25 mbar, <span class="html-italic">P</span> = 3 kW, <math display="inline"> <semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics> </math> = 50 sccm, CH<sub>4</sub> admixture: 2.5%, CO<sub>2</sub> admixture: 1% [<a href="#B66-photonics-03-00045" class="html-bibr">66</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">Figure 37
<p>Boltzmann plot of the P and R branch of CO lines of the first hot band for two pressure values, <span class="html-italic">p</span> = 0.25 and 0.35 mbar, <span class="html-italic">P</span> = 3 kW, <math display="inline"> <semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics> </math> = 50 sccm, CH<sub>4</sub> admixture: 2.5%, CO<sub>2</sub> admixture: 1% [<a href="#B66-photonics-03-00045" class="html-bibr">66</a>], © IOP Publishing. Reproduced with permission. All rights reserved.</p>
Full article ">
524 KiB  
Article
Development of a Multi-Objective Evolutionary Algorithm for Strain-Enhanced Quantum Cascade Lasers
by David Mueller and Gregory Triplett
Photonics 2016, 3(3), 44; https://doi.org/10.3390/photonics3030044 - 22 Jul 2016
Cited by 7 | Viewed by 4508
Abstract
An automated design approach using an evolutionary algorithm for the development of quantum cascade lasers (QCLs) is presented. Our algorithmic approach merges computational intelligence techniques with the physics of device structures, representing a design methodology that reduces experimental effort and costs. The algorithm [...] Read more.
An automated design approach using an evolutionary algorithm for the development of quantum cascade lasers (QCLs) is presented. Our algorithmic approach merges computational intelligence techniques with the physics of device structures, representing a design methodology that reduces experimental effort and costs. The algorithm was developed to produce QCLs with a three-well, diagonal-transition active region and a five-well injector region. Specifically, we applied this technique to Al x Ga 1 - x As/In y Ga 1 - y As strained active region designs. The algorithmic approach is a non-dominated sorting method using four aggregate objectives: target wavelength, population inversion via longitudinal-optical (LO) phonon extraction, injector level coupling, and an optical gain metric. Analysis indicates that the most plausible device candidates are a result of the optical gain metric and a total aggregate of all objectives. However, design limitations exist in many of the resulting candidates, indicating need for additional objective criteria and parameter limits to improve the application of this and other evolutionary algorithm methods. Full article
(This article belongs to the Special Issue Quantum Cascade Lasers - Advances and New Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>General flow diagram of the evolutionary algorithm used in this study.</p>
Full article ">Figure 2
<p>Generic gene sequence with value ranges.</p>
Full article ">Figure 3
<p>Schematic flow for the initial population.</p>
Full article ">Figure 4
<p>Al<math display="inline"> <semantics> <msub> <mrow/> <mi>x</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>x</mi> </mrow> </msub> </semantics> </math>As/In<math display="inline"> <semantics> <msub> <mrow/> <mi>y</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>y</mi> </mrow> </msub> </semantics> </math>As QC laser reference design, adopted from [<a href="#B27-photonics-03-00044" class="html-bibr">27</a>] with alloy proportions changed to <math display="inline"> <semantics> <mrow> <mi>x</mi> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>y</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics> </math> and simulated on GaAs (111).</p>
Full article ">Figure 5
<p>Plots of normalized average fitness per generation for (<b>a</b>) wavelength, (<b>b</b>) LO phonon, (<b>c</b>) injector coupling, and (<b>d</b>) gain objectives when initial population was generated from a training set.</p>
Full article ">Figure 6
<p>Plots of normalized average fitness per generation for (<b>a</b>) wavelength, (<b>b</b>) LO phonon, (<b>c</b>) injector coupling, and (<b>d</b>) gain objectives when initial population was generated randomly.</p>
Full article ">Figure 7
<p>Al<math display="inline"> <semantics> <msub> <mrow/> <mi>x</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>x</mi> </mrow> </msub> </semantics> </math>As/In<math display="inline"> <semantics> <msub> <mrow/> <mi>y</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>y</mi> </mrow> </msub> </semantics> </math>As QC laser candidate simulated on GaAs (111) with top aggregate fitness rank given a training initial population. In this design, <math display="inline"> <semantics> <mrow> <mi>x</mi> <mo>=</mo> <mn>0.21</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>y</mi> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics> </math> and the layer sequence (in nm) starting with the first injector well is: 3.0, <b>2.5</b>, 3.0, <b>1.8</b>, 2.0, <b>6.6</b>, 1.1, <b>0.8</b>, 3.0, <b>6.4</b>, <span class="html-italic">1.4,</span> <span class="html-italic"><b>1.1</b></span><span class="html-italic">, 4.5,</span> <span class="html-italic"><b>2.3</b></span><span class="html-italic">, 3.0,</span> <span class="html-italic"><b>2.8</b></span>. Barriers are in bold and the active region is shown in italics.</p>
Full article ">Figure 8
<p>Al<math display="inline"> <semantics> <msub> <mrow/> <mi>x</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>x</mi> </mrow> </msub> </semantics> </math>As/In<math display="inline"> <semantics> <msub> <mrow/> <mi>y</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>y</mi> </mrow> </msub> </semantics> </math>As QC laser candidate simulated on GaAs (111) with top objective (1) fitness rank given a training initial population. Relevant wave functions, <math display="inline"> <semantics> <msub> <mi>E</mi> <mn>3</mn> </msub> </semantics> </math> and <math display="inline"> <semantics> <msub> <mi>E</mi> <mn>2</mn> </msub> </semantics> </math> are in bold and colored.</p>
Full article ">Figure 9
<p>Al<math display="inline"> <semantics> <msub> <mrow/> <mi>x</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>x</mi> </mrow> </msub> </semantics> </math>As/In<math display="inline"> <semantics> <msub> <mrow/> <mi>y</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>y</mi> </mrow> </msub> </semantics> </math>As QC laser candidate simulated on GaAs (111) with top objective (2) fitness rank given a training initial population. Relevant wave functions, <math display="inline"> <semantics> <msub> <mi>E</mi> <mn>2</mn> </msub> </semantics> </math> and <math display="inline"> <semantics> <msub> <mi>E</mi> <mn>1</mn> </msub> </semantics> </math> are in bold and colored.</p>
Full article ">Figure 10
<p>Al<math display="inline"> <semantics> <msub> <mrow/> <mi>x</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>x</mi> </mrow> </msub> </semantics> </math>As/In<math display="inline"> <semantics> <msub> <mrow/> <mi>y</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>y</mi> </mrow> </msub> </semantics> </math>As QC laser candidate simulated on GaAs (111) with top objective (3) fitness rank given a training initial population. Relevant wave functions, <math display="inline"> <semantics> <msub> <mi>E</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>j</mi> <mo>,</mo> <mi>i</mi> <mi>j</mi> </mrow> </msub> </semantics> </math> are in bold and colored.</p>
Full article ">Figure 11
<p>Al<math display="inline"> <semantics> <msub> <mrow/> <mi>x</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>x</mi> </mrow> </msub> </semantics> </math>As/In<math display="inline"> <semantics> <msub> <mrow/> <mi>y</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>y</mi> </mrow> </msub> </semantics> </math>As QC laser candidate simulated on GaAs (111) with top objective (4) fitness rank given a training initial population.</p>
Full article ">Figure 12
<p>Al<math display="inline"> <semantics> <msub> <mrow/> <mi>x</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>x</mi> </mrow> </msub> </semantics> </math>As/In<math display="inline"> <semantics> <msub> <mrow/> <mi>y</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>y</mi> </mrow> </msub> </semantics> </math>As QC laser candidates simulated on GaAs (111). (<b>a</b>) candidate with top aggregate fitness rank and (<b>b</b>) candidate showing diversity of designs.</p>
Full article ">Figure 13
<p>Al<math display="inline"> <semantics> <msub> <mrow/> <mi>x</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>x</mi> </mrow> </msub> </semantics> </math>As/In<math display="inline"> <semantics> <msub> <mrow/> <mi>y</mi> </msub> </semantics> </math>Ga<math display="inline"> <semantics> <msub> <mrow/> <mrow> <mn>1</mn> <mo>-</mo> <mi>y</mi> </mrow> </msub> </semantics> </math>As QC laser candidates simulated on GaAs (111) with top fitness rank for (<b>a</b>) wavelength, (<b>b</b>) LO phonon, (<b>c</b>) injector coupling, and (<b>d</b>) gain objectives given a random initial population.</p>
Full article ">
2544 KiB  
Article
Fractional Effective Charges and Misner-Wheeler Charge without Charge Effect in Metamaterials
by Igor Smolyaninov
Photonics 2016, 3(3), 43; https://doi.org/10.3390/photonics3030043 - 8 Jul 2016
Cited by 1 | Viewed by 3936
Abstract
Transformation optics enables engineering of the effective topology and dimensionality of the optical space in metamaterials. Nonlinear optics of such metamaterials may mimic Kaluza-Klein theories having one or more kinds of effective charges. As a result, novel photon blockade devices may be realized. [...] Read more.
Transformation optics enables engineering of the effective topology and dimensionality of the optical space in metamaterials. Nonlinear optics of such metamaterials may mimic Kaluza-Klein theories having one or more kinds of effective charges. As a result, novel photon blockade devices may be realized. Here we demonstrate that an electromagnetic wormhole may be designed, which connects two points of such an optical space and changes its effective topological connectivity. Electromagnetic field configurations, which exhibit fractional effective charges, appear as a result of such topology change. Moreover, such effects as Misner-Wheeler “charge without charge” may be replicated. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) “Optical space” in an anisotropic uniaxial metamaterial may mimic such topologically non-trivial 3D space as R<sub>2</sub> × S<sub>1</sub>, which is a product of a 2D plane R<sub>2</sub> and a circle S<sub>1</sub>. (<b>b</b>) Schematic view of the “layered” 3D hyperbolic metamaterial made of subwavelength metal and dielectric layers, which can be used to emulate the R<sub>2</sub> × S<sub>1</sub> space. (<b>c</b>) Example of numerical calculations of electromagnetic field distribution in the <span class="html-italic">xz</span> plane inside the R<sub>2</sub> × S<sub>1</sub> metamaterial space using Comsol Multiphysics 4.2a (COMSOL, Inc., Burlington, MA, USA). The distances in the <span class="html-italic">xz</span> plane are measured in the units of wavelength. The power flow distribution (calculated using the scattering boundary conditions) is similar to power flow in a multimode planar waveguide.</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic view of a toroidal handlebody, which connects two points of the R<sub>2</sub> × S<sub>1</sub> metamaterial space and changes its effective topology. (<b>b</b>,<b>c</b>) Numerical simulations of the electromagnetic wormhole performed using the scattering boundary conditions: (b) spatial distribution of <span class="html-italic">ε<sub>1</sub></span>, and (c) corresponding spatial distribution of <span class="html-italic">B<sub>y</sub></span>. The distances in the <span class="html-italic">xz</span> plane are measured in the units of wavelength.</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>) Schematic view of a toroidal handlebody, which connects two points of the R<sub>2</sub> × S<sub>1</sub> metamaterial space and changes its effective topology. (<b>b</b>,<b>c</b>) Numerical simulations of the electromagnetic wormhole performed using the scattering boundary conditions: (b) spatial distribution of <span class="html-italic">ε<sub>1</sub></span>, and (c) corresponding spatial distribution of <span class="html-italic">B<sub>y</sub></span>. The distances in the <span class="html-italic">xz</span> plane are measured in the units of wavelength.</p>
Full article ">Figure 3
<p>Numerical simulations of power flow near the electromagnetic wormhole in the same geometry as in <a href="#photonics-03-00043-f002" class="html-fig">Figure 2</a>b,c, which demonstrates “charge without charge” effect. The average Poynting vector direction is indicated by black arrows. Near the wormhole openings the <span class="html-italic">z</span> components of the Poynting vector <span class="html-italic">S<sub>3</sub></span> are nonzero and have opposite signs, which according to Equations (11)–(14) leads to appearance of opposite contributions to the effective charge of the wormhole openings (the openings are marked by dashed lines).</p>
Full article ">Figure 4
<p>If the wormhole geometry exhibits some symmetry (such as symmetry under rotation by either π (<b>a</b>) or 2π/3 (<b>b</b>) radian, the allowed values of effective charges of the wormhole openings may become restricted to some simple fractions of the “elementary charge” <span class="html-italic">e</span>: if such a rotational symmetry is imposed on electromagnetic field configurations, the integer effective charge inside <span class="html-italic">A</span> must be divided equally between the wormhole openings.</p>
Full article ">Figure 4 Cont.
<p>If the wormhole geometry exhibits some symmetry (such as symmetry under rotation by either π (<b>a</b>) or 2π/3 (<b>b</b>) radian, the allowed values of effective charges of the wormhole openings may become restricted to some simple fractions of the “elementary charge” <span class="html-italic">e</span>: if such a rotational symmetry is imposed on electromagnetic field configurations, the integer effective charge inside <span class="html-italic">A</span> must be divided equally between the wormhole openings.</p>
Full article ">
2379 KiB  
Review
Recent Advances in Room Temperature, High-Power Terahertz Quantum Cascade Laser Sources Based on Difference-Frequency Generation
by Quanyong Lu and Manijeh Razeghi
Photonics 2016, 3(3), 42; https://doi.org/10.3390/photonics3030042 - 7 Jul 2016
Cited by 40 | Viewed by 6964
Abstract
We present the current status of high-performance, compact, THz sources based on intracavity nonlinear frequency generation in mid-infrared quantum cascade lasers. Significant performance improvements of our THz sources in the power and wall plug efficiency are achieved by systematic optimizing the device’s active [...] Read more.
We present the current status of high-performance, compact, THz sources based on intracavity nonlinear frequency generation in mid-infrared quantum cascade lasers. Significant performance improvements of our THz sources in the power and wall plug efficiency are achieved by systematic optimizing the device’s active region, waveguide, and chip bonding strategy. High THz power up to 1.9 mW and 0.014 mW for pulsed mode and continuous wave operations at room temperature are demonstrated, respectively. Even higher power and efficiency are envisioned based on enhancements in outcoupling efficiency and mid-IR performance. Our compact THz device with high power and wide tuning range is highly suitable for imaging, sensing, spectroscopy, medical diagnosis, and many other applications. Full article
(This article belongs to the Special Issue Quantum Cascade Lasers - Advances and New Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic of the DFG process between the electron states in a band structure of quantum cascade laser.</p>
Full article ">Figure 2
<p>(<b>a</b>) Band structure of a SPR structure at λ ~ 9 µm with large nonlinear DFG susceptibility based on a lattice-matched Ga<sub>0.47</sub>In<sub>0.53</sub>As/Al<sub>0.48</sub>In<sub>0.52</sub>As material system; and (<b>b</b>) band structure of a strongly-coupled SPR structure at λ ~ 7.8 µm based on a strain-balanced Al<sub>0.63</sub>In<sub>0.37</sub>As/Ga<sub>0.35</sub>In<sub>0.65</sub>As/Ga<sub>0.47</sub>In<sub>0.53</sub>As material system.</p>
Full article ">Figure 3
<p>(<b>a</b>) SEM images and (<b>b</b>) the Fourier analysis of the composite DFB gratings with 1 and 4 THz frequency spacing. The white bar in (<b>a</b>) corresponds to 4 µm.</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic of Čerenkov phase matching in a mid-IR quantum cascade lasers; (<b>b</b>) The thermal distributions of the epi-up and epi-down mounted Čerenkov devices; and (<b>c</b>,<b>d</b>) schematic of the epi-down mounting scheme and the SEM image of an epi-down mounted device.</p>
Full article ">Figure 5
<p>(<b>a</b>) Recently-demonstrated THz power records of the THz sources based on DFG QCLs at room temperature. THz power (<b>b</b>); spectra (<b>c</b>); and far field (<b>d</b>), as functions of currents for a high-power THz source based on DFG QCLs.</p>
Full article ">Figure 6
<p>SEM images of a buried ridge (<b>a</b>) and buried composite DFB grating waveguide (<b>b</b>); the white bars in (<b>a</b>) and (<b>b</b>) correspond to 10 and 2 µm, respectively; (<b>c</b>) <span class="html-italic">P-I-V</span> of a composite DFB THz device at room-temperature CW operation. Inset: CW lasing mid-IR spectrum at 1.62 A; and (<b>d</b>) THz CW power as a function of current. Inset: CW THz spectrum at 1.62 A.</p>
Full article ">
1730 KiB  
Article
Engineering Multi-Section Quantum Cascade Lasers for Broadband Tuning
by Steven Slivken and Manijeh Razeghi
Photonics 2016, 3(3), 41; https://doi.org/10.3390/photonics3030041 - 27 Jun 2016
Cited by 5 | Viewed by 5669
Abstract
In an effort to overcome current limitations to electrical tuning of quantum cascade lasers, a strategy is proposed which combines heterogeneous quantum cascade laser gain engineering with sampled grating architectures. This approach seeks to not only widen the accessible spectral range for an [...] Read more.
In an effort to overcome current limitations to electrical tuning of quantum cascade lasers, a strategy is proposed which combines heterogeneous quantum cascade laser gain engineering with sampled grating architectures. This approach seeks to not only widen the accessible spectral range for an individual emitter, but also compensate for functional non-uniformity of reflectivity and gain lineshapes. A trial laser with a dual wavelength core is presented which exhibits electroluminescence over a 750 cm−1 range and discrete single mode laser emission over a 700 cm−1 range. Electrical tuning over 180 cm−1 is demonstrated with a simple sampled grating design. A path forward to even wider tuning is also described using more sophisticated gain and grating design principles. Full article
(This article belongs to the Special Issue Quantum Cascade Lasers - Advances and New Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic evolution of original sampled grating distributed feedback (SGDFB) quantum cascade laser (QCL) to a broadband dual core SGDFB.</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic layer structure of dual core mid-wavelength infrared (MWIR) QCL; (<b>b</b>) Right axis: Normalized electroluminescence and Fabry Perot laser emission from the dual core wafer. Left axis: Threshold current density of DFB lasers according to targeted emission wavenumber. Points at the top left represent DFB lasers which exhibit only multimode emission.</p>
Full article ">Figure 3
<p>(<b>a</b>) Top view schematic of two section laser with sampled grating design specifics. Λ<sub>g</sub> = grating period. Λ<sub>s</sub> = sampling period (<b>b</b>) SEM cross-section of a buried ridge SGDFB laser. (<b>c</b>) Tuning characteristic of a dual core SGDFB laser.</p>
Full article ">Figure 4
<p>Evolution of electrically tunable QCL design for extreme spectral coverage.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop