[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (78)

Search Parameters:
Keywords = pressure sensitive paint

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 12503 KiB  
Article
Screen-Printed PVDF Piezoelectric Pressure Transducer for Unsteadiness Study of Oblique Shock Wave Boundary Layer Interaction
by Bei Wang, Cosimo Corsi, Thomas Weiland, Zhenyu Wang, Thomas Grund, Olaf Pohl, Johannes Max Bienia, Julien Weiss and Ha Duong Ngo
Micromachines 2024, 15(12), 1423; https://doi.org/10.3390/mi15121423 - 27 Nov 2024
Viewed by 752
Abstract
Shock wave boundary/layer interactions (SWBLIs) are critical in high-speed aerodynamic flows, particularly within supersonic regimes, where unsteady dynamics can induce structural fatigue and degrade vehicle performance. Conventional measurement techniques, such as pressure-sensitive paint (PSP), face limitations in frequency response, calibration complexity, and intrusive [...] Read more.
Shock wave boundary/layer interactions (SWBLIs) are critical in high-speed aerodynamic flows, particularly within supersonic regimes, where unsteady dynamics can induce structural fatigue and degrade vehicle performance. Conventional measurement techniques, such as pressure-sensitive paint (PSP), face limitations in frequency response, calibration complexity, and intrusive instrumentation. Similarly, MEMS-based sensors, like Kulite® sensors, present challenges in terms of intrusiveness, cost, and integration complexity. This study presents a flexible, lightweight polyvinylidene fluoride (PVDF) piezoelectric sensor array designed for high-resolution wall-pressure measurements in SWBLI research. The primary objective is to optimize low-frequency pressure fluctuation detection, addressing SWBLI’s need for accurate, real-time measurements of low-frequency unsteadiness. Fabricated using a double-sided screen-printing technique, this sensor array is low-cost, flexible, and provides stable, high-sensitivity data. Finite Element Method (FEM) simulations indicate that the sensor structure also has potential for high-frequency responses, behaving as a high-pass filter with minimal signal attenuation up to 300 kHz, although the current study’s experimental testing is focused on low-frequency calibration and validation. A custom low-frequency sound pressure setup was used to calibrate the PVDF sensor array, ensuring uniform pressure distribution across sensor elements. Wind tunnel tests at Mach 2 verified the PVDF sensor’s ability to capture pressure fluctuations and unsteady behaviors consistent with those recorded by Kulite sensors. The findings suggest that PVDF sensors are promising alternatives for capturing low-frequency disturbances and intricate flow structures in advanced aerodynamic research, with high-frequency performance to be further explored in future work. Full article
(This article belongs to the Special Issue MEMS/NEMS Devices and Applications, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Schlieren image of the flow field of the shock wave/boundary layer interaction [<a href="#B3-micromachines-15-01423" class="html-bibr">3</a>].</p>
Full article ">Figure 2
<p>Schematic of the design of sensor elements.</p>
Full article ">Figure 3
<p>(<b>a</b>) Displacement (mm) of PVDF film caused by 1 kPa pressure vertically to the left sensor electrode at 1 kHz, and (<b>b</b>) the electric potential caused by the piezoelectric effect of the PVDF film.</p>
Full article ">Figure 4
<p>Defined profile lines on the top of the Ag electrodes printed on PVDF film.</p>
Full article ">Figure 5
<p>Simulation of signal crosstalk between two adjacent sensor elements at varying distances <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>W</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>(<b>a</b>) The displacement (mm) on the profile lines of FEM simulation in frequency domain analysis, and (<b>b</b>) the electric potential caused by the displacement in frequency domain analysis.</p>
Full article ">Figure 7
<p>Spectra representation of displacement (mm) and electric potential (V) in frequency domain up to 300 kHz.</p>
Full article ">Figure 8
<p>Illustration of 3D-integrated thin-film flexible PVDF sensor array: (<b>a</b>–<b>c</b>) show the process steps.</p>
Full article ">Figure 9
<p>Images of PVDF sensor array with 4 mm sensor elements.</p>
Full article ">Figure 10
<p>(<b>a</b>) The charge amplifier circuit for the PVDF piezoelectric sensor array, and (<b>b</b>) the band-pass behavior of the charge amplifier circuit [<a href="#B31-micromachines-15-01423" class="html-bibr">31</a>,<a href="#B34-micromachines-15-01423" class="html-bibr">34</a>].</p>
Full article ">Figure 11
<p>Block diagram of calibration setup.</p>
Full article ">Figure 12
<p>Setup of the sound pressure calibration test equipment (<b>a</b>) and the PVDF sensor array test samples (<b>b</b>).</p>
Full article ">Figure 13
<p>Geometric modeling of the sound pressure calibration test setup (<b>a</b>), and the pressure contour distribution (<b>b</b>).</p>
Full article ">Figure 14
<p>Schematic presentation of the defined profile line in acoustic simulation (<b>a</b>), and (<b>b</b>) the total sound pressure level (dB) on the profile line at frequency sweep.</p>
Full article ">Figure 15
<p>(<b>a</b>) Heterogeneous and (<b>b</b>) homogeneous sensor arrays.</p>
Full article ">Figure 16
<p>Sensor response of the heterogeneous PVDF sensor array in time domain.</p>
Full article ">Figure 17
<p>(<b>a</b>) Amplitude responses of sensors of different sizes as the driven voltage from the acoustic wave generator increases, and (<b>b</b>) sensitivities of each sensor element in the heterogeneous array at a constant pressure of <math display="inline"><semantics> <mrow> <mi>P</mi> <mo>=</mo> <mn>240</mn> <mo> </mo> <mi>Pa</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 18
<p>(<b>a</b>) The sensitivity of PVDF sensors (4 mm) and the Kulite sensor vs. increasing pressure, and (<b>b</b>) the sensor sensitivity of each sensor element of the test homogeneous PVDF sensor array at pressure <math display="inline"><semantics> <mrow> <mi>P</mi> <mo>=</mo> <mn>240</mn> <mo> </mo> <mi>Pa</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 19
<p>Schematic representation of the supersonic wind tunnel present at the Technical University of Berlin with a focus on the turbulent SBLI test section [<a href="#B28-micromachines-15-01423" class="html-bibr">28</a>].</p>
Full article ">Figure 20
<p>Installation of pressure sensor on plug (<b>a</b>), and (<b>b</b>) the plug was embedded in the test section in the wind tunnel.</p>
Full article ">Figure 21
<p>Schematic of (<b>a</b>) shock wave/boundary layer interaction with Mach 2.0 flow intake in 3d-presentation, and (<b>b</b>) the sensor mounting in the measurement.</p>
Full article ">Figure 22
<p>PVDF sensor responses at 6 positions in time domain.</p>
Full article ">Figure 23
<p>Value <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>·</mo> <mi>P</mi> <mi>S</mi> <mi>D</mi> </mrow> </semantics></math> of 6 PVDF sensors S1 to S6 in frequency domain.</p>
Full article ">Figure 24
<p>Schematic of the flow field and respective <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>·</mo> <mi>P</mi> <mi>S</mi> <mi>D</mi> </mrow> </semantics></math> spectra along the streamwise direction for the PVDF sensor array.</p>
Full article ">Figure 25
<p>Comparison of values <math display="inline"><semantics> <mrow> <mi mathvariant="normal">f</mi> <mo>·</mo> <mrow> <mi>PSD</mi> <mtext> </mtext> <mi>vs</mi> </mrow> <mo>.</mo> <mrow> <mtext> </mtext> <mi>Strouhal</mi> </mrow> </mrow> </semantics></math> number <math display="inline"><semantics> <mrow> <msub> <mi>S</mi> <mi>t</mi> </msub> </mrow> </semantics></math> at six investigation positions ① to ⑥ with Kulite sensors under the same conditions in the wind tunnel.</p>
Full article ">Figure 26
<p>Illustrates the similarity of the pressure signal values received by the two types of sensors at the same location and time.</p>
Full article ">
14 pages, 3436 KiB  
Article
Advancing Sustainability: Geraniol-Enhanced Waterborne Acrylic Pressure-Sensitive Adhesives without Chemical Modification
by Ludovica Di Lorenzo, Simone Bordignon, Michele R. Chierotti, Ignazio Andrea Alfeo, Adrian Krzysztof Antosik and Valentina Brunella
Materials 2024, 17(20), 4957; https://doi.org/10.3390/ma17204957 - 10 Oct 2024
Viewed by 1230
Abstract
The escalating global emphasis on sustainability, coupled with stringent regulatory frameworks, has spurred the quest for environmentally viable alternatives to petroleum-derived materials. Within this context, the adhesives industry has been actively seeking renewable options and eco-friendly synthesis pathways. This study introduces geraniol, a [...] Read more.
The escalating global emphasis on sustainability, coupled with stringent regulatory frameworks, has spurred the quest for environmentally viable alternatives to petroleum-derived materials. Within this context, the adhesives industry has been actively seeking renewable options and eco-friendly synthesis pathways. This study introduces geraniol, a monoterpenoid alcohol, in its unmodified form, as a key component in the production of waterborne pressure-sensitive adhesives (PSAs) based on acrylic latex through emulsion polymerization. Multiple formulations were developed at varying reaction times. The adhesives underwent comprehensive chemical characterization employing techniques such as Fourier-transform infrared spectroscopy (FTIR), thermogravimetric analysis (TGA), differential scanning calorimetry (DSC), Nuclear Magnetic Resonance (NMR), Gel Permeation Chromatography (GPC), and dynamic light scattering (DLS). The viscosities of the formulations were measured between 4000 and 5000 cP. Adhesion tests showed peel strength values of 0.52 N/mm on cardboard and 0.32 N/mm on painted steel for the geraniol-based formulations. The results demonstrate the potential for geraniol-based PSAs to offer a sustainable alternative to petroleum-derived adhesives, with promising thermal and adhesive properties. Full article
(This article belongs to the Section Polymeric Materials)
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>H (600.17 MHz) solution NMR spectra of PSA blank (<b>top</b>), *PSA 3 (<b>center</b>), and physical mixture (<b>bottom</b>) acquired in CDCl<sub>3</sub> at room temperature. The green dashed box highlights the range of <sup>1</sup>H chemical shifts in which we would see resonances ascribable to unreacted geraniol, if present.</p>
Full article ">Figure 2
<p><sup>13</sup>C (150.91 MHz) solution NMR spectra of PSA blank (<b>top</b>), *PSA 3 (<b>center</b>), and physical mixture (<b>bottom</b>), acquired in CDCl<sub>3</sub> at room temperature. The green dashed box highlights the range of <sup>13</sup>C chemical shifts in which we would see resonances ascribable to unreacted geraniol, if present.</p>
Full article ">Figure 3
<p>FTIR spectra of reference PSA blank and geraniol-based PSAs.</p>
Full article ">Figure 4
<p>Zoom in the 1900–1500 [1/cm] range of the FTIR spectra of geraniol-based PSAs compared with the FTIR spectra of the main starting monomers (<b>a</b>) butyl acrylate and (<b>b</b>) geraniol.</p>
Full article ">Figure 5
<p>(<b>a</b>) TGA for acrylic latex PSA with different reaction times (PSA blank 2.5 h; PSA 1_4h; PSA 2_5 h; PSA 3_6 h). (<b>b</b>) TGA of PSA 3 and *PSA 3 with post-polymerization treatment.</p>
Full article ">Figure 6
<p>A 180° peeling test of the acrylic latex PSAs.</p>
Full article ">
25 pages, 8196 KiB  
Article
Temperature-Dependent Viscosity Analysis of Powell–Eyring Fluid Model during a Roll-over Web Coating Process
by Fateh Ali, Srikantha Narasimhamurthy, Soniya Hegde and Muhammad Usman
Polymers 2024, 16(12), 1723; https://doi.org/10.3390/polym16121723 - 17 Jun 2024
Viewed by 1070
Abstract
The roll coating method is of considerable significance in several industries, as it is applied practically in the production of paint, the manufacturing of PVC-coated cloth, and the plastic industry. The current study theoretically and computationally analyses the Powell–Eyring fluids with variable viscosity [...] Read more.
The roll coating method is of considerable significance in several industries, as it is applied practically in the production of paint, the manufacturing of PVC-coated cloth, and the plastic industry. The current study theoretically and computationally analyses the Powell–Eyring fluids with variable viscosity during the non-isothermal roll-over web phenomenon. Based on the lubrication approximation theory (LAT), the problem was formulated. The system of partial differential equations (PDEs) obtained from the mathematical modeling was further simplified to a set of ordinary differential equations (ODEs) using suitable transformations. A regular perturbation method was implemented to obtain the solution in terms of velocity, pressure gradient, pressure, and flow rate per unit width. This study also captures important engineering characteristics such as coating thickness, Nusselt number, shear stress, roll/sheet separating force, and roll-transmitted power to the fluid. Along with a comparison between the present work and published work, both graphical and tabular representations wer made to study the effects of various factors. It was observed that the velocity profile is the decreasing function of non-Newtonian and Reynold viscosity parameters. In addition, the response surface methodology (RSM) was employed to investigate the sensitivity of the shear stress and the Nusselt number. Full article
Show Figures

Figure 1

Figure 1
<p>Geometry for rollover web coating process.</p>
Full article ">Figure 2
<p>Impact of <math display="inline"><semantics> <mrow> <mi>W</mi> <mi>e</mi> </mrow> </semantics></math> on velocity profile.</p>
Full article ">Figure 3
<p>Impact of <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> </mrow> </semantics></math> on velocity profile.</p>
Full article ">Figure 4
<p>Impact of <math display="inline"><semantics> <mi>m</mi> </semantics></math> on velocity profile.</p>
Full article ">Figure 5
<p>Impact of <math display="inline"><semantics> <mi>N</mi> </semantics></math> on velocity profile.</p>
Full article ">Figure 6
<p>Impact of <math display="inline"><semantics> <mrow> <mi>W</mi> <mi>e</mi> </mrow> </semantics></math> on P. G.</p>
Full article ">Figure 7
<p>Impact of <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> </mrow> </semantics></math> on P. G.</p>
Full article ">Figure 8
<p>Impact of <math display="inline"><semantics> <mi>m</mi> </semantics></math> on P. G.</p>
Full article ">Figure 9
<p>Impact of <math display="inline"><semantics> <mi>N</mi> </semantics></math> on P. G.</p>
Full article ">Figure 10
<p>Impact of <math display="inline"><semantics> <mrow> <mi>W</mi> <mi>e</mi> </mrow> </semantics></math> on pressure profile.</p>
Full article ">Figure 11
<p>Impact of <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> </mrow> </semantics></math> on the pressure profile.</p>
Full article ">Figure 12
<p>Impact of <math display="inline"><semantics> <mi>m</mi> </semantics></math> on pressure profile.</p>
Full article ">Figure 13
<p>Impact of <math display="inline"><semantics> <mi>N</mi> </semantics></math> on the pressure profile.</p>
Full article ">Figure 14
<p>Impact of <math display="inline"><semantics> <mi>m</mi> </semantics></math> on the Nusselt number.</p>
Full article ">Figure 15
<p>Streamline pattern and surface plots for <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>W</mi> <mi>e</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>=</mo> <mn>1.3015</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>0.00233</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 16
<p>Contour plots and 3D plots for (<b>a</b>) <math display="inline"><semantics> <mi>N</mi> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mi>N</mi> </semantics></math> and <math display="inline"><semantics> <mi>m</mi> </semantics></math> (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mi>m</mi> </semantics></math>, respectively, and the third variable is maintained at an intermediate level throughout on <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 16 Cont.
<p>Contour plots and 3D plots for (<b>a</b>) <math display="inline"><semantics> <mi>N</mi> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mi>N</mi> </semantics></math> and <math display="inline"><semantics> <mi>m</mi> </semantics></math> (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mi>m</mi> </semantics></math>, respectively, and the third variable is maintained at an intermediate level throughout on <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 17
<p>Contour plots and 3D plots for (<b>a</b>) <math display="inline"><semantics> <mi>N</mi> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mi>N</mi> </semantics></math> and <math display="inline"><semantics> <mi>m</mi> </semantics></math> (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mi>m</mi> </semantics></math>, respectively, and the third variable is maintained at an intermediate level throughout on <math display="inline"><semantics> <mrow> <mi>N</mi> <mi>u</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 17 Cont.
<p>Contour plots and 3D plots for (<b>a</b>) <math display="inline"><semantics> <mi>N</mi> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mi>N</mi> </semantics></math> and <math display="inline"><semantics> <mi>m</mi> </semantics></math> (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>r</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mi>m</mi> </semantics></math>, respectively, and the third variable is maintained at an intermediate level throughout on <math display="inline"><semantics> <mrow> <mi>N</mi> <mi>u</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 18
<p>Sensitivity of response variable <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </msub> </mrow> </semantics></math> when (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>=</mo> <mo>-</mo> <mn>1</mn> <mo>,</mo> <mi>B</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mi>B</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>B</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 19
<p>Sensitivity of response variable <math display="inline"><semantics> <mrow> <mi>N</mi> <mi>u</mi> </mrow> </semantics></math> when (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>=</mo> <mo>-</mo> <mn>1</mn> <mo>,</mo> <mi>C</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mi>C</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>C</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 20
<p>(<b>a</b>–<b>c</b>): Optimized <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>N</mi> <mi>u</mi> </mrow> </semantics></math> solution with desirability factors.</p>
Full article ">
17 pages, 5484 KiB  
Article
Comparison of Lifetime-Based Pressure-Sensitive Paint Measurements in a Wind Tunnel Using Model Pitch–Traverse and Pitch–Pause Modes
by Christian Klein, Daisuke Yorita and Ulrich Henne
Photonics 2024, 11(6), 546; https://doi.org/10.3390/photonics11060546 - 7 Jun 2024
Viewed by 1156
Abstract
In order to improve the data productivity of a wind tunnel test, the model under investigation in the wind tunnel is moved continuously with a predetermined constant angular speed in the so-called pitch–traverse mode. Alternatively, the wind tunnel model can be moved in [...] Read more.
In order to improve the data productivity of a wind tunnel test, the model under investigation in the wind tunnel is moved continuously with a predetermined constant angular speed in the so-called pitch–traverse mode. Alternatively, the wind tunnel model can be moved in the so-called pitch–pause mode, in which it keeps its position for a certain (measurement) time at a fixed pitch position, after which it is moved to the next pitch position. The latter procedure is more time-consuming, so, for the same time interval, the number of measured data points taken in the pitch–pause mode is less than that for the pitch–traverse mode. Since wind tunnel test time can be quite expensive, in most wind tunnel tests where only conventional forces and pressures are recorded with conventional measuring systems, the wind tunnel model is moved in the pitch–traverse mode in order to obtain as much aerodynamic data as possible during the tunnel runtime. The application of the Pressure-Sensitive Paint (PSP) technique has been widely used in wind tunnel testing for the purpose of providing pressure data on wind tunnel models with high spatial resolution. The lifetime-based PSP method has several advantages over the intensity-based method since it often has higher accuracy. Up until now, the lifetime-based PSP technique has mainly been used for wind tunnel testing, where the test model has been moved to the pitch–pause mode. The traditional lifetime method using on-chip accumulation requires multiple (~1000) excitation light pulses to accumulate enough luminescence (fluorescence or phosphorescence) photons on the camera sensor to provide acceptable signal-to-noise ratios and, therefore, it may seem to be not compatible with a continuously moving wind tunnel model. Nevertheless, the present study verifies the application of lifetime-based PSP utilizing on-chip accumulation with a continuously moving wind tunnel model which would make the entire PSP data acquisition compatible with that of the conventional measurements (forces and pressures), as mentioned above. In this paper, the applicability of the lifetime-based PSP technique to a continuously moving wind tunnel model (in pitch–traverse mode) is investigated with the help of measurements in the transonic wind tunnel in Göttingen (TWG). For this investigation, PSP was applied on the delta-wing model DLR-F22, which is to be tested in TWG. The pressure distribution on the wind tunnel model was measured using the PSP lifetime method for both model movement modes (pitch–pause and pitch–traverse mode) so that the corresponding PSP results could be directly compared with each other. In addition, an error analysis of the PSP results was carried out and compared with the conventional pressure measurement results, hence providing an assessment of the accuracy of the PSP results; finally, a recommendation for future PSP measurements could be given. Full article
(This article belongs to the Special Issue Editorial Board Members’ Collection Series: Photonics Sensors)
Show Figures

Figure 1

Figure 1
<p>PSP coating layers.</p>
Full article ">Figure 2
<p>(<b>a</b>) Luminescence lifetime decay curves for different pressure levels at <span class="html-italic">T</span> = 30 °C and camera gate settings. (<b>b</b>) Pressure sensitivity curves for different temperatures of the lifetime-based method.</p>
Full article ">Figure 3
<p>(<b>a</b>) Geometry of DLR-F22 model [<a href="#B4-photonics-11-00546" class="html-bibr">4</a>]. (<b>b</b>) Photo of the installed DLR-F22 model in the perforated wall test section of TWG. The area illuminated by the LEDs on the wind tunnel model is shown in magenta color.</p>
Full article ">Figure 4
<p>Model coating inside the portable painting booth.</p>
Full article ">Figure 5
<p>PSP image acquisition sequences in pitch–pause mode. (<b>a</b>) At one test point (e.g., at one angle-of-attack <span class="html-italic">α</span>), it takes 13 s to acquire the one pressure image. (<b>b</b>) Example of a pitch–pause run at three different <span class="html-italic">α</span> = 16, 20, 24°.</p>
Full article ">Figure 6
<p>(<b>a</b>) Example of a pitch–traverse sweep from <span class="html-italic">α</span> = 15°→25°, the total time for pitch–traverse was 100 s. Zoom shows the image acquisition in the first two seconds. (<b>b</b>) Image acquisition sequence during a pitch–traverse sweep.</p>
Full article ">Figure 7
<p>Flow chart of the PSP data processing.</p>
Full article ">Figure 8
<p>Pressure distribution measured by means of lifetime PSP in pitch–pause mode on the DLR-F22 model, M = 0.85, <span class="html-italic">α</span> = 16°, <span class="html-italic">β</span> = 0° with vortex (black) and shock (white) positions indicated by dashed/dotted lines.</p>
Full article ">Figure 9
<p>PSP result images obtained in pitch–pause (<b>a</b>) or pitch–traverse mode (<b>b</b>) for M = 0.85, <span class="html-italic">α</span> = 16°, <span class="html-italic">β</span> = 0°; black and white lines showing vortex and shock locations in (<b>a</b>) have been transposed directly into (<b>b</b>) to facilitate comparison.</p>
Full article ">Figure 10
<p><span class="html-italic">C<sub>p</sub></span>-difference image (<b>a</b>) and RMS<sub><span class="html-italic">Gate</span>2</sub>-result image (<b>b</b>) for M = 0.85, <span class="html-italic">α</span> = 16°, <span class="html-italic">β</span> = 0° (see text).</p>
Full article ">Figure 11
<p>PSP result images obtained in pitch–pause (<b>a</b>) or pitch–traverse mode (<b>b</b>). (<b>c</b>) <span class="html-italic">C<sub>p</sub></span>-difference image for M = 0.95, <span class="html-italic">α</span> = 16°, <span class="html-italic">β</span> = 0°.</p>
Full article ">Figure 12
<p>PSP result images obtained in pitch–pause (<b>a</b>) or pitch–traverse mode (<b>b</b>) for M = 1.1, <span class="html-italic">α</span> = 16°, <span class="html-italic">β</span> = 0°. (<b>c</b>) <span class="html-italic">C<sub>p</sub></span>-difference image—see text.</p>
Full article ">Figure 13
<p>Error analysis for all 150 data points measured with PSP for M = 0.85 in pitch–traverse mode. (<b>a</b>) Comparison of PSP pressure values and the corresponding Kulite<sup>®</sup> results, (<b>b</b>) pressure difference (PSP-to-Kulite<sup>®</sup>).</p>
Full article ">Figure A1
<p>Pressure distribution for M = 0.5, <span class="html-italic">α</span> = 16°, <span class="html-italic">β</span> = 0° measured in pitch–pause mode.</p>
Full article ">
16 pages, 7843 KiB  
Article
Characterization of the Endwall Flow in a Low-Pressure Turbine Cascade Perturbed by Periodically Incoming Wakes, Part 2: Unsteady Blade Surface Measurements Using Pressure-Sensitive Paint
by Tobias Schubert, Dragan Kožulović and Martin Bitter
Aerospace 2024, 11(5), 404; https://doi.org/10.3390/aerospace11050404 - 16 May 2024
Cited by 1 | Viewed by 1190
Abstract
Unsteady pressure-sensitive paint (i-PSP) measurements were performed at a sampling rate of 30 kHz to investigate the near-endwall blade suction surface flow inside a low-pressure turbine cascade operating at engine-relevant high-speed and low-Re conditions. The investigation focuses on the interaction of periodically incoming [...] Read more.
Unsteady pressure-sensitive paint (i-PSP) measurements were performed at a sampling rate of 30 kHz to investigate the near-endwall blade suction surface flow inside a low-pressure turbine cascade operating at engine-relevant high-speed and low-Re conditions. The investigation focuses on the interaction of periodically incoming bar wakes at 500 Hz with the secondary flow and the blade suction surface. The results build on extensive PIV measurements presented in the first part of this two-part publication, which captured the ’negative-jet-effect’ of the wakes throughout the blade passage. The surface pressure distributions are combined with CFD to analyze the flow topology, such as the passage vortex separation line. By analyzing data from phase-locked PIV and PSP measurements, a wake-induced moving pressure gradient negative in space and positive in time is found, which is intensified in the secondary flow region by 33% with respect to midspan. Furthermore, two methods of frequency-filtering based on FFT and SPOD are compared and utilized to associate a pressure fluctuation peak around 678 Hz with separation bubble oscillation. Full article
(This article belongs to the Special Issue Advanced Flow Diagnostic Tools)
Show Figures

Figure 1

Figure 1
<p>Illustration of the T106A test case featuring a split flat plate endwall (yellow) and moving bars upstream of the low-pressure turbine blades (cyan); adapted from Schubert et al. [<a href="#B8-aerospace-11-00404" class="html-bibr">8</a>].</p>
Full article ">Figure 2
<p>Simulated entropy generation rate (<math display="inline"><semantics> <msubsup> <mover accent="true"> <mi>S</mi> <mo>˙</mo> </mover> <mi>v</mi> <mo>*</mo> </msubsup> </semantics></math> non-dimensionalized) at several axial slices inside the T106A blade passage under steady (<b>a</b>) and periodically unsteady inflow conditions (<b>b</b>); adapted from Schubert et al. [<a href="#B13-aerospace-11-00404" class="html-bibr">13</a>]; (HSVp—pressure side leg of horseshoe vortex; PV—passage vortex; CRV—counter-rotating vortex; CV—corner vortex; LE—leading edge; TE—trailing edge).</p>
Full article ">Figure 3
<p>Experimental setup for unsteady pressure-sensitive paint (i-PSP) measurements in the T106A turbine cascade; (<b>a</b>) schematic of the cross-sectional side view; (<b>b</b>) camera access via a mirror and UV-light illumination; (<b>c</b>) camera view on the suction surface of the measurement blade (#4).</p>
Full article ">Figure 4
<p>Comparison of the surface pressure measured with PSP, to static pressure taps at the midspan of the T106A turbine cascade under periodically unsteady inflow conditions; (<b>a</b>) relative pressure difference; (<b>b</b>) isentropic Mach number distribution.</p>
Full article ">Figure 5
<p>Phase-locking and synchronization method; (<b>a</b>–<b>c</b>) alignment of the spatial bar position with the temporal trigger signals of the bar counter; (<b>d</b>) assignment of a timestamp <math display="inline"><semantics> <mi>τ</mi> </semantics></math> based on the rel. position in a bar passing period; (<b>e</b>) phase-averaging by binning in <math display="inline"><semantics> <mi>τ</mi> </semantics></math>-bins of 25 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>s; (<b>f</b>) application of the synchronization process to CFD by setting the reference bar position to <math display="inline"><semantics> <mi>τ</mi> </semantics></math> = 0.</p>
Full article ">Figure 6
<p>Measured static pressure distribution and supplemental CFD on the T106A suction surface under steady (<b>a</b>–<b>c</b>) and periodically unsteady (<b>d</b>–<b>f</b>) inflow conditions; (<b>a</b>,<b>d</b>) time-averaged pressure (PSP); (<b>b</b>,<b>e</b>) standard deviation of pressure fluctuations (PSP); (<b>c</b>,<b>f</b>) wall shear stress (CFD).</p>
Full article ">Figure 7
<p>Power spectrum density of the mean pressure amplitudes (PSP) at midspan under periodically unsteady inflow conditions.</p>
Full article ">Figure 8
<p>PSP measurements on the T106A suction surface stimulated by the first harmonic wake generator frequency of 502 Hz and the separation bubble osculation at 678 Hz; (<b>a</b>) FFT-based filtered pressure amplitudes; (<b>b</b>,<b>c</b>) first SPOD mode of the measured intensity ratios averaged over 100 time steps.</p>
Full article ">Figure 9
<p>Phase-locked PSP pressure fluctuations measured on the T106A suction surface at four time steps of the bar passing period (the respective wake position is indicate above and the mean flow direction is from left to right).</p>
Full article ">Figure 10
<p>Phase-locked PIV measurements of the bar wake velocity deficit in the T106A turbine passage close to the endwall (<math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> </mrow> </semantics></math> = 0.2) at two time steps and its effect on the blade surface pressure; adapted from part 1 [<a href="#B14-aerospace-11-00404" class="html-bibr">14</a>].</p>
Full article ">Figure 11
<p>Time-space correlation of the measured static pressure distributions at midspan of the suction surface (<b>a</b>,<b>b</b>) and close to the endwall at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> </mrow> </semantics></math> = 0.2 (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 12
<p>Comparison of CFD predictions of pressure fluctuations to phase-locked PSP measurements at position <math display="inline"><semantics> <mrow> <mi>P</mi> <mn>1</mn> </mrow> </semantics></math> (<b>left</b>) &amp; <math display="inline"><semantics> <mrow> <mi>P</mi> <mn>2</mn> </mrow> </semantics></math> (<b>right</b>) marked in <a href="#aerospace-11-00404-f009" class="html-fig">Figure 9</a>.</p>
Full article ">
16 pages, 13924 KiB  
Article
Characterization of the Endwall Flow in a Low-Pressure Turbine Cascade Perturbed by Periodically Incoming Wakes, Part 1: Flow Field Investigations with Phase-Locked Particle Image Velocimetry
by Tobias Schubert, Dragan Kožulović and Martin Bitter
Aerospace 2024, 11(5), 403; https://doi.org/10.3390/aerospace11050403 - 16 May 2024
Cited by 1 | Viewed by 1080
Abstract
Particle image velocimetry (PIV) measurements were performed inside a low-pressure turbine cascade operating at engine-relevant high-speed and low-Re conditions to investigate the near-endwall flow. Of particular research interest was the dominant periodic disturbance of the flow field by incoming wakes, which were generated [...] Read more.
Particle image velocimetry (PIV) measurements were performed inside a low-pressure turbine cascade operating at engine-relevant high-speed and low-Re conditions to investigate the near-endwall flow. Of particular research interest was the dominant periodic disturbance of the flow field by incoming wakes, which were generated by moving cylindrical bars at a frequency of 500 Hz. Two PIV setups were utilized to resolve both (1) a large blade-to-blade plane close to the endwall as well as midspan and (2) the wake effects in an axial flow field downstream of the blade passage. The measurements were performed using a phase-locked approach in order to align and compare the results with comprehensive CFD data that are also available for this test case. The experimental results not only support a better understanding and even a quantification of the wake-induced over/under-turning inside and downstream of the passage, they also enable the tracing of the ‘negative-jet-effect’, which is widely known in the CFD branch of the turbomachinery community but is seldom visualized in experiments. The results also reveal that the bar wake periodically widens the blade wake by up to 165%, while the secondary flow is less affected and exhibits a phase lag with respect to the 2D-flow effects. The results presented here are an essential basis for the subsequent investigation of the near-endwall blade suction surface effects using unsteady pressure-sensitive paint in the second part of this two-part publication. Full article
(This article belongs to the Special Issue Advanced Flow Diagnostic Tools)
Show Figures

Figure 1

Figure 1
<p>Illustration of the T106A test case featuring a split flat plate endwall (yellow) and moving bars upstream of the low-pressure turbine blades (cyan); adapted from Schubert et al. [<a href="#B8-aerospace-11-00403" class="html-bibr">8</a>].</p>
Full article ">Figure 2
<p>Simulated entropy generation rate (<math display="inline"><semantics> <msubsup> <mover accent="true"> <mi>S</mi> <mo>˙</mo> </mover> <mi>v</mi> <mo>*</mo> </msubsup> </semantics></math> non-dimensionalized) at several axial slices inside the T106A blade passage under steady (<b>a</b>) and periodically unsteady inflow conditions (<b>b</b>); adapted from Schubert et al. [<a href="#B13-aerospace-11-00403" class="html-bibr">13</a>]; (HSVp—pressure side leg of horseshoe vortex; PV—passage vortex; CRV—counter-rotating vortex; CV—corner vortex; LE—leading edge; TE—trailing edge).</p>
Full article ">Figure 3
<p>The 2D2C PIV setup for blade-to-blade measurements inside the T106A linear cascade; (<b>a</b>) schematic of the cross-sectional side view; (<b>b</b>) endoscopic camera setup; (<b>c</b>) optical access into the flow channel; (<b>d</b>) PIV laser sheet at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> </mrow> </semantics></math> = 0.2; (<b>e</b>) camera calibration target in the field of view.</p>
Full article ">Figure 4
<p>PIV setup for 2D3C measurements in an axial plane (MP 2) located <math display="inline"><semantics> <mrow> <mn>40</mn> <mo>%</mo> <mspace width="0.166667em"/> <msub> <mi>C</mi> <mi>x</mi> </msub> </mrow> </semantics></math> downstream of the blade passage; (<b>a</b>) schematic of the cross-sectional side view; (<b>b</b>) camera setup and laser guidance to the test section; (<b>c</b>) laser sheet in the cascade exit flow.</p>
Full article ">Figure 5
<p>Phase-locking and synchronization method; (<b>a</b>–<b>c</b>) alignment of the spatial bar position with the temporal trigger signals of the bar counter; (<b>d</b>) assignment of a timestamp <math display="inline"><semantics> <mi>τ</mi> </semantics></math> based on the relative position during a bar passing period; (<b>e</b>) binning and subsequent averaging in <math display="inline"><semantics> <mi>τ</mi> </semantics></math>-bins of 0.1 ms; (<b>f</b>) application of the synchronization process to CFD by setting the reference bar position to <math display="inline"><semantics> <mi>τ</mi> </semantics></math> = 0.</p>
Full article ">Figure 6
<p>Time-averaged velocity distributions illustrated by isentropic Mach number in the downstream flow field (MP 2) under steady (<b>a</b>–<b>c</b>) and periodically unsteady inflow conditions (<b>d</b>–<b>f</b>).</p>
Full article ">Figure 7
<p>Time-averaged passage flow field at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> </mrow> </semantics></math> = 0.2 (near-endwall) under steady (<b>left</b>) and periodically unsteady (<b>middle</b>) inflow conditions; (<b>right</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>|</mo> <mi>V</mi> <mo>|</mo> </mrow> </semantics></math> between the two cases of inflow conditions.</p>
Full article ">Figure 8
<p>Time-averaged flow angle difference with respect to midspan at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> </mrow> </semantics></math> = 0.2 (<math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>β</mi> <mrow> <mi>s</mi> <mi>e</mi> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mi>β</mi> <mo>−</mo> <msub> <mover accent="true"> <mi>β</mi> <mo>¯</mo> </mover> <mrow> <mi>M</mi> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math>) under steady (<b>left</b>) and periodically unsteady (<b>middle</b>) inflow conditions; (<b>right</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>β</mi> <mrow> <mi>s</mi> <mi>e</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> measured with a five-hole-probe at <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <msub> <mi>C</mi> <mi>x</mi> </msub> <mo>=</mo> <mn>1.4</mn> </mrow> </semantics></math> downstream of the cascade.</p>
Full article ">Figure 9
<p>Transport of the bar wake velocity deficit through the T106A turbine passage close to the endwall (<math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> </mrow> </semantics></math> = 0.2) at four time steps and its effect on the downstream flow field.</p>
Full article ">Figure 10
<p>Evolution of the bar wake effect on the near-endwall (<math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> </mrow> </semantics></math> = 0.2) passage flow field at position <math display="inline"><semantics> <mrow> <mi>P</mi> <mn>1</mn> </mrow> </semantics></math> (see <a href="#aerospace-11-00403-f010" class="html-fig">Figure 10</a>) over time; comparison of velocity deficit and the flow angle between CFD and phase-locked PIV (gray positions in the time plot mark the displays in <a href="#aerospace-11-00403-f009" class="html-fig">Figure 9</a>).</p>
Full article ">Figure 11
<p>Periodic effect of the bar wakes on the downstream flow field (MP 2); (<b>a</b>) secondary flow extension <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>y</mi> <mrow> <mi>r</mi> <mi>e</mi> <mi>l</mi> </mrow> </msub> <mo>=</mo> <mrow> <mo>(</mo> <mi>y</mi> <mo>−</mo> <mfenced open="&#x2329;" close="&#x232A;"> <mi>y</mi> </mfenced> <mo>)</mo> </mrow> <mo>/</mo> <mi>P</mi> </mrow> </semantics></math> and intensity <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>T</mi> <mi>K</mi> <msub> <mi>E</mi> <mrow> <mi>r</mi> <mi>e</mi> <mi>l</mi> </mrow> </msub> <mo>=</mo> <mrow> <mo>(</mo> <mi>T</mi> <mi>K</mi> <mi>E</mi> <mo>−</mo> <mfenced separators="" open="&#x2329;" close="&#x232A;"> <mi>T</mi> <mi>K</mi> <mi>E</mi> </mfenced> <mo>)</mo> </mrow> <mo>/</mo> <mfenced separators="" open="&#x2329;" close="&#x232A;"> <mi>T</mi> <mi>K</mi> <mi>E</mi> </mfenced> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> </mrow> </semantics></math> = 0.2; (<b>b</b>) blade wake extension <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>y</mi> <mrow> <mi>r</mi> <mi>e</mi> <mi>l</mi> </mrow> </msub> </mrow> </semantics></math> at midspan.</p>
Full article ">Figure 12
<p>Illustration of the experimental data set available for the T106A test case under unsteady inflow conditions; (<b>left</b>) duplicated PIV passage flow fields presented in this paper and time-averaged isentropic surface Mach number on the blade suction surface measured with unsteady pressure-sensitive paint; (<b>right</b>) high-resolution surface pressure amplitude measurement (PSP) stimulated by the dominant wake generator frequency around 500 Hz. Black dots show markers for image mapping.</p>
Full article ">
12 pages, 4797 KiB  
Article
Global Surface Pressure Pattern for a Compressible Elliptical Cavity Flow Using Pressure-Sensitive Paint
by Yi-Xuan Huang and Kung-Ming Chung
Aerospace 2024, 11(2), 159; https://doi.org/10.3390/aerospace11020159 - 15 Feb 2024
Viewed by 1230
Abstract
The flow field in a cavity depends on the properties of the upstream boundary layer and the cavity geometry. Comprehensive studies for rectangular cavities have been conducted. This experimental study determines the global surface pressure pattern for elliptical cavities (eccentricities of 0, 0.66 [...] Read more.
The flow field in a cavity depends on the properties of the upstream boundary layer and the cavity geometry. Comprehensive studies for rectangular cavities have been conducted. This experimental study determines the global surface pressure pattern for elliptical cavities (eccentricities of 0, 0.66 and 0.87) in a naturally developed turbulent boundary layer using pressure-sensitive paint. The ratio between the length (major axis) and the depth is 4.43–21.5, and the freestream Mach number is 0.83. The mean surface pressure distribution of an elliptical cavity resembles that of a rectangular cavity. A change in the value of eccentricity (wall curvature) affects the region for an adverse pressure gradient in an open cavity, an extension of the plateau in a transitional–closed cavity and flow expansion near the front and rear edges. The boundaries between an open, transitional and closed cavities vary. Full article
Show Figures

Figure 1

Figure 1
<p>Model setup.</p>
Full article ">Figure 2
<p>Calibration curve for PSP.</p>
Full article ">Figure 3
<p>Surface pressure distribution for a cavity with <span class="html-italic">L/H</span> = 4.43.</p>
Full article ">Figure 4
<p>Surface pressure distribution for a cavity with <span class="html-italic">L/H</span> = 6.14.</p>
Full article ">Figure 5
<p>Surface pressure distribution for a cavity with <span class="html-italic">L/H</span> = 14.33.</p>
Full article ">Figure 6
<p>Surface pressure distribution for a cavity with <span class="html-italic">L/H</span> = 21.50.</p>
Full article ">Figure 7
<p>Mean surface pressure distribution for <span class="html-italic">ε</span> = 0 (red solid line: PSP data; black points: Kulite data).</p>
Full article ">Figure 8
<p>Mean surface pressure distribution for <span class="html-italic">ε</span> = 0.66 (red solid line: PSP data; black points: Kulite data).</p>
Full article ">Figure 9
<p>Mean surface pressure distribution for <span class="html-italic">ε</span> = 0.87 (red solid line: PSP data; black points: Kulite data).</p>
Full article ">Figure 10
<p>The effect of eccentricity on rear-edge expansion.</p>
Full article ">
19 pages, 9453 KiB  
Article
Experimental and Numerical Investigations into the Effects of Rim Seal Structure on Endwall Film Cooling and Flow Field Characteristics
by Yixuan Lu, Zhao Liu, Weixin Zhang, Yuqiang Ding and Zhenping Feng
Energies 2023, 16(24), 7976; https://doi.org/10.3390/en16247976 - 8 Dec 2023
Cited by 1 | Viewed by 1080
Abstract
During the practical operation of gas turbines, relatively cooled air from the compressor and the rim seal is applied in order to prevent mainstream ingestion into the space between the rotor and stator disc cavities, which can prolong the service life of hot [...] Read more.
During the practical operation of gas turbines, relatively cooled air from the compressor and the rim seal is applied in order to prevent mainstream ingestion into the space between the rotor and stator disc cavities, which can prolong the service life of hot components. On the one hand, the purge flow from the rim seal will inevitably interact with the mainstream and result in secondary flow on the endwall. On the other hand, it can also provide an additional cooling effect. In this paper, four rim seal structures, including an original single-tooth seal (ORI), a double-tooth seal (DS), a single-tooth seal with an adverse direction of the coolant purge flow and mainstream (AS) and a double-tooth seal with an adverse direction of the coolant purge flow and mainstream (ASDS), are experimentally and numerically investigated with mass flow ratios of 0.5%, 1.0% and 1.5%. The flow orientation of the coolant from the rim seal is considered as one of the main factors. The pressure-sensitive paint technique is used to experimentally measure the film cooling effectiveness on the endwall, and flow field analysis is conducted via numerical simulations. The results show that the cooling effect decreases in the cases of DS and ASDS. AS and ASDS can achieve a better film cooling performance, especially under a higher mass flow ratio. Furthermore, the structural changes in the rim seal have little impact on the aerodynamic performance. AS and ASDS can both achieve a better aerodynamic and film cooling performance. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of experimental system: (<b>a</b>) 3D diagram of experimental system; (<b>b</b>) local diagram of the experimental section.</p>
Full article ">Figure 2
<p>Diagram and photo of test section.</p>
Full article ">Figure 3
<p>Distribution of the pressure coefficient.</p>
Full article ">Figure 4
<p>Graphic model of the seal geometry.</p>
Full article ">Figure 5
<p>Graphic model and partial diagram of different seal geometry structures: (<b>a</b>) ORI; (<b>b</b>) DS; (<b>c</b>) AS; (<b>d</b>) ASDS.</p>
Full article ">Figure 6
<p>Photograph of the PSP measurement process from different views: (<b>a</b>) overall view of the experiment; (<b>b</b>) partial view of the experimental cascade.</p>
Full article ">Figure 7
<p>Velocity distribution of the mainstream.</p>
Full article ">Figure 8
<p>Validation of the turbulence model: (<b>a</b>) contours of <span class="html-italic">η</span>; (<b>b</b>) laterally averaged <span class="html-italic">η</span>.</p>
Full article ">Figure 9
<p>Local refined regions of the grid.</p>
Full article ">Figure 10
<p>Grid independence analysis.</p>
Full article ">Figure 11
<p>Experimental contours <span class="html-italic">η</span> of four seal structures with different <span class="html-italic">MFR</span>s: (<b>a</b>) ORI; (<b>b</b>) DS; (<b>c</b>) AS; (<b>d</b>) ASDS.</p>
Full article ">Figure 12
<p>The comparison of laterally averaged <span class="html-italic">η</span> of four seal structures with different <span class="html-italic">MFR</span>s: (<b>a</b>) <span class="html-italic">MFR</span> = 0.5%; (<b>b</b>) <span class="html-italic">MFR</span> = 1.0%; (<b>c</b>) <span class="html-italic">MFR</span> = 1.5%.</p>
Full article ">Figure 13
<p>The comparison of area-averaged <span class="html-italic">η</span> of four seal structures with different <span class="html-italic">MFR</span>s.</p>
Full article ">Figure 14
<p>The cross-sections of rim seal.</p>
Full article ">Figure 15
<p>Streamline distribution of ORI with different <span class="html-italic">MFR</span>s: (<b>a</b>) <span class="html-italic">MFR</span> = 0.5%; (<b>b</b>) <span class="html-italic">MFR</span> = 1.0%; (<b>c</b>) <span class="html-italic">MFR</span> = 1.5%.</p>
Full article ">Figure 16
<p>Streamline distribution of DS with different <span class="html-italic">MFR</span>s: (<b>a</b>) <span class="html-italic">MFR</span> = 0.5%; (<b>b</b>) <span class="html-italic">MFR</span> = 1.0%; (<b>c</b>) <span class="html-italic">MFR</span> = 1.5%.</p>
Full article ">Figure 17
<p>Streamline distribution of AS with different <span class="html-italic">MFR</span>s: (<b>a</b>) <span class="html-italic">MFR</span> = 0.5%; (<b>b</b>) <span class="html-italic">MFR</span> = 1.0%; (<b>c</b>) <span class="html-italic">MFR</span> = 1.5%.</p>
Full article ">Figure 17 Cont.
<p>Streamline distribution of AS with different <span class="html-italic">MFR</span>s: (<b>a</b>) <span class="html-italic">MFR</span> = 0.5%; (<b>b</b>) <span class="html-italic">MFR</span> = 1.0%; (<b>c</b>) <span class="html-italic">MFR</span> = 1.5%.</p>
Full article ">Figure 18
<p>Streamline distribution of ASDS with different <span class="html-italic">MFR</span>s: (<b>a</b>) <span class="html-italic">MFR</span> = 0.5%; (<b>b</b>) <span class="html-italic">MFR</span> = 1.0%; (<b>c</b>) <span class="html-italic">MFR</span> = 1.5%.</p>
Full article ">Figure 19
<p>ζ of four seal structures with different <span class="html-italic">MFR</span>s: (<b>a</b>) ORI; (<b>b</b>) DS; (<b>c</b>) AS; (<b>d</b>) ASDS.</p>
Full article ">Figure 20
<p>Area-averaged <span class="html-italic">ζ</span> of four seal structures with different <span class="html-italic">MFR</span>s.</p>
Full article ">
17 pages, 4478 KiB  
Article
Temperature Dependency Model in Pressure Measurement for the Motion-Capturing Pressure-Sensitive Paint Method
by Daiki Kurihara and Hirotaka Sakaue
Sensors 2023, 23(24), 9714; https://doi.org/10.3390/s23249714 - 8 Dec 2023
Viewed by 1064
Abstract
Pressure-sensitive paint (PSP) has received significant attention for capturing surface pressure in recent years. One major source of uncertainty in PSP measurements, temperature dependency, stems from the fundamental photophysical process that allows PSP to extract pressure information. The motion-capturing PSP method, which involves [...] Read more.
Pressure-sensitive paint (PSP) has received significant attention for capturing surface pressure in recent years. One major source of uncertainty in PSP measurements, temperature dependency, stems from the fundamental photophysical process that allows PSP to extract pressure information. The motion-capturing PSP method, which involves two luminophores, is introduced as a method to reduce the measurement uncertainty due to temperature dependency. A theoretical model for the pressure uncertainty due to temperature dependency is proposed and demonstrated using a static pressure measurement with an applied temperature gradient. The experimental validation of the proposed model shows that the motion-capturing PSP method reduces the temperature dependency by 37.7% compared to the conventional PSP method. The proposed model also proves that a PSP with zero temperature dependency is theoretically possible. Full article
(This article belongs to the Special Issue Optical Sensors for Flow Diagnostics II)
Show Figures

Figure 1

Figure 1
<p>Conceptual description of the temperature dependency for the motion-capturing PSP method.</p>
Full article ">Figure 2
<p>Illustrative example of pressure uncertainty due to temperature for considered cases.</p>
Full article ">Figure 3
<p>Schematic of the experimental validation setup.</p>
Full article ">Figure 4
<p>In situ temperature calibration.</p>
Full article ">Figure 5
<p>Pressure uncertainty due to temperature distribution based on (<b>a</b>) the motion-capturing PSP method and (<b>b</b>) the intensity method, and (<b>c</b>) the temperature profile along the line A–B.</p>
Full article ">Figure 6
<p>Pressure uncertainty profiles due to temperature based on the motion-capturing PSP method in blue and the intensity method in red. Dashed lines show the model prediction. The blue and red areas are model uncertainties.</p>
Full article ">Figure A1
<p>Pressure calibration setup.</p>
Full article ">Figure A2
<p>Stern–Volmer plot.</p>
Full article ">Figure A3
<p>Thermocouple temperature profiles.</p>
Full article ">Figure A4
<p>Spectra of the two-color PSP varying with pressure.</p>
Full article ">
14 pages, 8420 KiB  
Article
Experimental Study on the Improvement of the Film Cooling Effectiveness of Various Modified Configurations Based on a Fan-Shaped Film Cooling Hole on a Flat Plate
by Seokmin Kim, DongEun Lee, Young Seok Kang and Dong-Ho Rhee
Energies 2023, 16(23), 7752; https://doi.org/10.3390/en16237752 - 24 Nov 2023
Cited by 1 | Viewed by 1093
Abstract
Modern gas turbines have evolved by increasing the turbine inlet temperature (TIT) to improve performance. This development has led to a demand for cooling techniques. Among these, the film cooling, which involves injecting compressed air through holes on the turbine surface, is a [...] Read more.
Modern gas turbines have evolved by increasing the turbine inlet temperature (TIT) to improve performance. This development has led to a demand for cooling techniques. Among these, the film cooling, which involves injecting compressed air through holes on the turbine surface, is a prominent cooling technique used to protect the turbine surface. In this study, a comparative analysis is conducted between the conventional fan-shaped film cooling hole, primarily used in film cooling techniques, and modified shapes achieved by altering the geometry of the film cooling hole based on a fan-shaped hole to assess and compare the cooling performance on a flat plate surface. The adiabatic film cooling effectiveness was measured for three film cooling holes, the Baseline of a 7-7-7 fan-shaped film cooling hole, namely, Staircase, which had a double-step at the hole exit, and Compound Expansion, which had an additional expanded flow path at the hole leading edge. The used measurement technique was the pressure-sensitive paint (PSP) technique, using nitrogen gas as the foreign gas, and experiments were conducted at a density ratio of 1.0 and blowing ratios ranging from 0.5 to 2.0. The results reveal that the modified holes featured wider lateral expansion at the hole exits, resulting in a broader distribution of the cooling effectiveness in the lateral direction compared to the Baseline. The Staircase shows a better performance, although an overall cooling effectiveness trend similar to that of the Baseline. Furthermore, the Compound Expansion demonstrates an enhancement in the cooling performance with an increased blowing ratio, notably achieving nearly double the cooling effectiveness compared to that of the Baseline at a blowing ratio of 2.0. Full article
(This article belongs to the Section J1: Heat and Mass Transfer)
Show Figures

Figure 1

Figure 1
<p>7-7-7 Fan-shaped film cooling hole configuration (Baseline) [<a href="#B20-energies-16-07752" class="html-bibr">20</a>].</p>
Full article ">Figure 2
<p>Baseline with a staircase geometry at the hole exit (Staircase) [<a href="#B20-energies-16-07752" class="html-bibr">20</a>].</p>
Full article ">Figure 3
<p>Baseline with an additional flow passage at the hole leading edge (Compound Expansion) [<a href="#B21-energies-16-07752" class="html-bibr">21</a>].</p>
Full article ">Figure 4
<p>PSP calibration curve.</p>
Full article ">Figure 5
<p>A schematic diagram of the experimental facility [<a href="#B20-energies-16-07752" class="html-bibr">20</a>].</p>
Full article ">Figure 6
<p>The test plate model adopted in this test (Baseline) [<a href="#B20-energies-16-07752" class="html-bibr">20</a>].</p>
Full article ">Figure 7
<p>Velocity and turbulence intensity profile at x/D = −12 [<a href="#B20-energies-16-07752" class="html-bibr">20</a>].</p>
Full article ">Figure 8
<p>Contour plots of the film cooling effectiveness for the various hole configurations at M = 0.5.</p>
Full article ">Figure 9
<p>Contour plots of the film cooling effectiveness for the various hole configurations at M = 1.0.</p>
Full article ">Figure 10
<p>Contour plots of the film cooling effectiveness for the various hole configurations at M = 2.0.</p>
Full article ">Figure 11
<p>Schematics of the flow structure around the Compound Expansion hole.</p>
Full article ">Figure 12
<p>Lateral distribution of the film cooling effectiveness at x/D = 2.</p>
Full article ">Figure 13
<p>Lateral distribution of the film cooling effectiveness at x/D = 5.</p>
Full article ">Figure 14
<p>Comparison of the laterally averaged film cooling effectiveness.</p>
Full article ">Figure 15
<p>Area-averaged film cooling effectiveness for the various hole configurations [<a href="#B26-energies-16-07752" class="html-bibr">26</a>].</p>
Full article ">
19 pages, 10878 KiB  
Article
Experimental Study on the Improvement of Film Cooling Effectiveness of Various Modified Configurations Based on a Fan-Shaped Film Cooling Hole on an Endwall
by Seokmin Kim, DongEun Lee, Young Seok Kang and Dong-Ho Rhee
Energies 2023, 16(23), 7733; https://doi.org/10.3390/en16237733 - 23 Nov 2023
Cited by 2 | Viewed by 1095
Abstract
Several studies have previously been conducted to improve the cooling performance of film cooling. However, most of the research has conducted experiments with film cooling holes on flat plates, and thus, the results of these studies do not encompass the influence of the [...] Read more.
Several studies have previously been conducted to improve the cooling performance of film cooling. However, most of the research has conducted experiments with film cooling holes on flat plates, and thus, the results of these studies do not encompass the influence of the complex mainstream behavior within the turbine passage on film cooling. In this study, three different film cooling hole configurations were installed on the endwall of a turbine linear cascade to measure adiabatic film cooling effectiveness and evaluate cooling performance. The film cooling holes compared in the experiment for film cooling effectiveness were a 7-7-7 fan-shaped hole (Baseline), a Baseline with a double-step structure at the hole exit (Staircase), and a Baseline with an additional expanded passage at the hole leading edge (Compound Expansion). A total of nine holes were manufactured on the turbine endwall to assess film cooling performance, as various factors, such as mainstream acceleration, secondary flow within the turbine passage, and so on, can influence film cooling. Adiabatic film cooling effectiveness was measured using the pressure-sensitive paint (PSP) technique. Mass flow ratios ranging from 0.25% to 1.25% of the mass flow rate of a single turbine passage were supplied to the plenum chamber within the test rig. As a result, all experimental results confirmed the impact of secondary flow within the turbine passage on film cooling. In the case of the Staircase, it exhibits an overall cooling trend similar to the Baseline. It shows small cooling performance degradation compared with Baseline due to lift-off, and its double-step structure laterally expanding results in better cooling performance at high mass flow ratio (MFR) conditions. For the Compound Expansion, at low MFR, the momentum of the coolant is lower compared with other configurations, leading to lower cooling performance due to the influence of secondary flow. However, at high MFR, the Compound Expansion provides wider protection compared with other hole geometries and shows high cooling performance. Full article
(This article belongs to the Section J1: Heat and Mass Transfer)
Show Figures

Figure 1

Figure 1
<p>Flow visualization of passage vortex and induced vortices [<a href="#B15-energies-16-07733" class="html-bibr">15</a>].</p>
Full article ">Figure 2
<p>Visualization of coolant trajectories on the endwall [<a href="#B16-energies-16-07733" class="html-bibr">16</a>].</p>
Full article ">Figure 3
<p>Configurations of test film cooling hole [<a href="#B22-energies-16-07733" class="html-bibr">22</a>]. (<b>a</b>) 7-7-7 Fan-shaped film cooling hole (Baseline); (<b>b</b>) Baseline with staircase geometry (Staircase); (<b>c</b>) Baseline with additional expanded flow passage (Compound Expansion).</p>
Full article ">Figure 4
<p>Turbine blade endwall specimen with hole locations [<a href="#B23-energies-16-07733" class="html-bibr">23</a>].</p>
Full article ">Figure 5
<p>Test plenum and specimen for turbine cascade experiment. (<b>a</b>) 3D CAD model [<a href="#B23-energies-16-07733" class="html-bibr">23</a>]; (<b>b</b>) Real test model.</p>
Full article ">Figure 6
<p>A schematic diagram of the blade endwall experimental facility [<a href="#B23-energies-16-07733" class="html-bibr">23</a>].</p>
Full article ">Figure 7
<p>PSP calibration curve.</p>
Full article ">Figure 8
<p>Example of the experiment using PSP. (<b>a</b>) The specimen coated with PSP; (<b>b</b>) Image during the experiment; (<b>c</b>) The experimental result after data processing.</p>
Full article ">Figure 9
<p>Test blade for velocity profile measurement and velocity profile at x/C = 0.2 from turbine blade leading edge. (<b>a</b>) Test blade for velocity profile measurement; (<b>b</b>) Velocity profile at x/C = 0.2.</p>
Full article ">Figure 10
<p>Contour plots of film cooling effectiveness for the Baseline.</p>
Full article ">Figure 11
<p>Contour plots of film cooling effectiveness for the Staircase.</p>
Full article ">Figure 12
<p>Contour plots of film cooling effectiveness for the Compound Expansion.</p>
Full article ">Figure 13
<p>Contour plots of film cooling effectiveness at the second row of holes at MFR = 0.50%.</p>
Full article ">Figure 14
<p>Contour plots of film cooling effectiveness at the second row of holes at MFR = 1.00%.</p>
Full article ">Figure 15
<p>Regions of area-averaged film cooling effectiveness.</p>
Full article ">Figure 16
<p>Area-averaged film cooling effectiveness in each region.</p>
Full article ">Figure 17
<p>Coverage ratio in each region.</p>
Full article ">Figure 18
<p>Contour plots of film cooling effectiveness at the Region 3.</p>
Full article ">Figure 19
<p>Area-averaged film cooling effectiveness for various hole configurations [<a href="#B22-energies-16-07733" class="html-bibr">22</a>].</p>
Full article ">Figure 20
<p>Area-averaged film cooling effectiveness and coverage ratio in entire region. (<b>a</b>) Area-averaged effectiveness; (<b>b</b>) Coverage ratio.</p>
Full article ">
22 pages, 9006 KiB  
Article
Fast-Responding Pressure-Sensitive Paint Measurements of the IC3X at Mach 7.2
by Valeria Delgado Elizondo, Abinayaa Dhanagopal and Christopher S. Combs
Aerospace 2023, 10(10), 890; https://doi.org/10.3390/aerospace10100890 - 18 Oct 2023
Cited by 2 | Viewed by 2417
Abstract
Global surface pressure measurements of a 5.7% scale AFRL Initial Concept 3.X vehicle (IC3X) were obtained using a fast-responding ruthenium-based pressure-sensitive paint (PSP) at the UTSA Mach 7 Ludwieg Tube Wind Tunnel at two different angles of attack, 0° and 2.5°. Static calibration [...] Read more.
Global surface pressure measurements of a 5.7% scale AFRL Initial Concept 3.X vehicle (IC3X) were obtained using a fast-responding ruthenium-based pressure-sensitive paint (PSP) at the UTSA Mach 7 Ludwieg Tube Wind Tunnel at two different angles of attack, 0° and 2.5°. Static calibration of the paint was performed over a range of 0.386 kPa to 82.7 kPa to relate luminescent intensity to pressure. Details on the facility, paint preparation, application, calibration, and image processing techniques are provided in the manuscript. The results from statistical, spectral, and proper orthogonal decomposition (POD) analyses are presented to characterize the pressure field observed on the model. The experimental results qualitatively follow the expected trends and correspond to the occurrence of shock waves and expansion fans, which were visualized via Schlieren imaging. The theoretical pressure range obtained from conical shock analysis for 0° agrees with the experimentally derived pressure range for the model, and the outliers are attributed to errors in image registration. This study presents preliminary pressure measurements that pave the way for obtaining time-resolved global PSP measurements to train and validate aerothermodynamic machine learning models. Full article
Show Figures

Figure 1

Figure 1
<p>CAD rendering of the UTSA Mach 7 Ludwieg Tube Wind Tunnel.</p>
Full article ">Figure 2
<p>CAD rendering of the UTSA IC3X wind tunnel model.</p>
Full article ">Figure 3
<p>IC3X PSP wind tunnel experimental setup.</p>
Full article ">Figure 4
<p>PSP static experimental setup block diagram.</p>
Full article ">Figure 5
<p>Data processing roadmap.</p>
Full article ">Figure 6
<p>Experimental pressure-sensitive paint static calibration curves.</p>
Full article ">Figure 7
<p>Coherence of paint response on a canonical flat plate model from [<a href="#B39-aerospace-10-00890" class="html-bibr">39</a>].</p>
Full article ">Figure 8
<p>PSP degradation rate estimates.</p>
Full article ">Figure 9
<p>Raw and corrected intensity at constant pressure.</p>
Full article ">Figure 10
<p>Overlay of Schlieren- and PSP-derived pressure field at 0° AoA.</p>
Full article ">Figure 11
<p>Instantaneous pressure fields.</p>
Full article ">Figure 12
<p>Average pressure fields.</p>
Full article ">Figure 13
<p>Pressure time histories and power spectral densities at various locations on the vehicle surface at 0°AoA under hypersonic flow.</p>
Full article ">Figure 13 Cont.
<p>Pressure time histories and power spectral densities at various locations on the vehicle surface at 0°AoA under hypersonic flow.</p>
Full article ">Figure 14
<p>Pressure time histories and power spectral densities at various locations on the vehicle surface at 2.5° AoA under hypersonic flow.</p>
Full article ">Figure 14 Cont.
<p>Pressure time histories and power spectral densities at various locations on the vehicle surface at 2.5° AoA under hypersonic flow.</p>
Full article ">Figure 15
<p>Pressure distribution at various locations on the model surface at (<b>a</b>) 2.5° AoA and (<b>b</b>) 0° AoA.</p>
Full article ">Figure 16
<p>Probability density function curves derived from PSP pressure measurements at various locations of interest at (<b>a</b>) 0° AoA and (<b>b</b>) 2.5° AoA under hypersonic flow.</p>
Full article ">Figure 17
<p>Results of 0° AoA mode 1-200.</p>
Full article ">Figure 18
<p>Results of 2.5° AoA Mode 1-200.</p>
Full article ">Figure 19
<p>POD mode energy vs. mode number.</p>
Full article ">Figure 20
<p>Optical sensors on the IC3X leading edge.</p>
Full article ">Figure 21
<p>Pressure time series of the IC3X model stagnation location at 0°AoA.</p>
Full article ">Figure 22
<p>Pressure error bar of the IC3X model stagnation location at 0°AoA.</p>
Full article ">Figure 23
<p>PDF of the IC3X model stagnation pressure at 0°AoA.</p>
Full article ">Figure 24
<p>Theoretical pressure on the IC3X model P<sub>2</sub> at M = 7.2 + 0.2, γ = 1.4 over half cone angle θ = 8°–20° at 0° AoA.</p>
Full article ">
19 pages, 5898 KiB  
Article
Innovation in Green Materials for the Non-Contact Stabilization of Sensitive Works of Art: Preliminary Assessment and the First Application of Ultra-Low Viscosity Hydroxypropyl Methylcellulose (HPMC) by Ultrasonic Misting to Consolidate Unstable Porous and Powdery Media
by Tomas Markevicius
Sustainability 2023, 15(20), 14699; https://doi.org/10.3390/su152014699 - 10 Oct 2023
Viewed by 4494
Abstract
Paintings and other works of art created with fragile and mechanically unstable powdery media present challenges to conservators. Frequently, powdery media is water-sensitive, extremely fragile, tends to delaminate, and may be altered by even the slightest physical action or interaction with liquids. Materials [...] Read more.
Paintings and other works of art created with fragile and mechanically unstable powdery media present challenges to conservators. Frequently, powdery media is water-sensitive, extremely fragile, tends to delaminate, and may be altered by even the slightest physical action or interaction with liquids. Materials that can provide an efficient stabilization without unacceptably altering the optical characteristics of the delicate substrate are extremely limited. Among these, Funori, Isinglass, and Methocel A4C have become established for this use. In bench practice, consolidants are frequently applied in a non-contact way, using ultrasonic and pneumatic aerosol generators to minimize the impact of the consolidant on sensitive substrates. However, nebulizing the available materials is problematic in bench practice, because of their high viscosity and, only extremely low concentrations can be nebulized using low kinetic impact ultrasonic or pressure-based misting systems adopted from the healthcare industry. As a potential innovative solution, this study introduces novel ultra-low viscosity (ULV) cellulose ethers (ULV-HPMC) for stabilisation of unstable porous and powdery surfaces, which have been successfully applied in bench practice for the pilot treatment of Edvard Munch painting on canvas and two 19th c. Thai gouache paintings on panel. Novel ULV-HPMC materials have multiple desirable qualities for consolidation treatments in conservation, and in accelerated aging tests marginally outperformed Methocel A4C, considered to be one of the most stable consolidants in the practice of conservation. Because of the ultra-low viscosity, higher concentrations of ULV-HPMC materials can be applied as water-based aerosols in a non-contact way and in fewer applications, which is a significant advantage in the treatment of delicate water-sensitive surfaces. Notably, novel ULV biopolymers are low-cost, derive from sustainable and renewable sources, and do not raise health and environmental concerns. Such novel materials and methods seamlessly resonate with the ICOM-CC’s Melbourne 2014 declaration, EU Green Deal, and the UN’s Sustainable Development goals and show potential adding new sustainable materials with the exceptionally low viscosity to the conservator’s tool box. Full article
Show Figures

Figure 1

Figure 1
<p>SEM image of natural ochre gouache paint on paper shows powdery media on the surface (top). Stereo microscope image of powdery delaminating paint of E. Munch Alma Mater figure study (1912) oil on canvas, M 881/Woll M 31, © Munch Museum, Oslo, Norway.</p>
Full article ">Figure 2
<p>The author in the process of consolidation using ULV-HPMC and ultrasonic misting under the stereomicroscope in the treatment of Edvard Munch’s unvarnished painting “Alma Mater” (1911–16), Munch Museum, Oslo. Paintings conservator Carolina Jimenez Gray in the process of the consolidation of unvarnished traditional Thai Thotsachat gouache paintings on panel using a pneumatic aerosol generator and HPMC, Victoria and Albert Museum, London.</p>
Full article ">Figure 3
<p>Chemical structure of hydroxypropyl methylcellulose C<sub>56</sub>H<sub>108</sub>O<sub>30.</sub></p>
Full article ">Figure 4
<p>MOBS 3P4 aerosol generation test using ultrasonic nebulizer AGS 2000.</p>
Full article ">Figure 5
<p>Colour change of the samples after aging (dark ambient and accelerated aging in climatic chamber).</p>
Full article ">Figure 6
<p>Image after aging: MC Methocel A4C, ULV-HPMC MOBS 6P4, Isinglass, Funori samples. Image: Lora Angelova.</p>
Full article ">Figure 7
<p>Gloss change (measured in gloss units (GU) along the vertical axis) of the samples after the application of the consolidation medium on paper before (blue) and after (green) aging (dark ambient—<b>left</b>; accelerated aging in climatic chamber—<b>right</b>).</p>
Full article ">Figure 8
<p>Stereo microscope image of cobalt blue pastel (not analysed) on 100% Cotton Archival Rag Endleaf paper after consolidation using MOBS 3P3 3.5 wt%. Image: Lucia Pereira Pardo.</p>
Full article ">Figure 9
<p>Methocel A4C, MOBS 3P4, and MOBS 6P4 were applied by brushing (2 wt%, 4 wt%) and misting (1.5 wt%, 3.5 wt%), A4C was only brush-applied (2 wt%, 4 wt%). Image: Lucia Pereira Pardo.</p>
Full article ">Figure 10
<p>Colour change of the mock-up paint samples after consolidation (B2: brushed 2 wt%; B4: brushed 4 wt%; M1.5: misted 1.5 wt%; M3.5: misted 3.5 wt%).</p>
Full article ">Figure 11
<p>Gloss change after consolidation (B2: brushed 2 wt%; B4: brushed 4 wt%; M1.5: misted 1.5 wt%; M3.5: misted 3.5 wt%).</p>
Full article ">Figure 12
<p>Results of the peeling test (mg of pigment removed): B2: brush applied 2 wt%; B4: brush applied 4 wt%, M1.5: misted 1.5 wt%, M3.5: misted 3.5 wt%.</p>
Full article ">Figure 13
<p>Consolidation tests on matte powdery natural yellow ochre and Isinglass glue tempera paint mock-up. Testing consolidation materials—a: Funori in water 3 wt% brushed, b: Isinglass in water 0.7 wt% brushed, c: MOBS 3P4 in water 3.5 wt% brushed, d: MOBS 3P4 3.5%/Isinglass 0.7 wt% in water, brushed, e: MOBS 3P4 3.5 wt%/Isinglass 0.7 wt% in water, misted, f: Lascaux<sup>®</sup> medium for consolidation in water, brushed, g: Aquazol-200 4 wt% in ethanol, brushed, h (marked white): MOBS 6P4 4 wt% in water, misted (result deemed optimal for the treatment).</p>
Full article ">Figure 14
<p>Edvard Munch Alma Mater, figure study (1912) oil on canvas M 881/Woll M 311 © Munch Museum. Edvard Munch “Alma Mater” (<b>left</b>), application of ULV-HPMC using ultrasonic misting on “Alma Mater” (<b>centre</b>) of the grid coordinate system, used to map the consolidation process (<b>right</b>).</p>
Full article ">Figure 15
<p>Thai Thotsachat, 57.4 × 46.7 cm, gouache and gilding on panel (IS.43-2005), Victoria &amp; Albert Museum, London. Overall view before stabilization treatment (<b>left</b>), a detail of upper right quadrant before (<b>centre</b>) and after consolidation (<b>right</b>). © Victoria and Albert Museum, London, United Kingdom. Images: Carolina Jimenez Gray.</p>
Full article ">
15 pages, 7945 KiB  
Article
The Development and Application of Two-Color Pressure-Sensitive Paint in Jet Impingement Experiments
by Wei-Chieh Chen, Chih-Yung Huang, Kui-Thong Tan and Hirotaka Sakaue
Aerospace 2023, 10(9), 805; https://doi.org/10.3390/aerospace10090805 - 15 Sep 2023
Cited by 1 | Viewed by 1342
Abstract
This study aimed to develop a two-color pressure-sensitive paint (PSP) that has both high pressure sensitivity and high temperature sensitivity. Different nitrobenzoxadiazole (NBD) derivatives were used as the temperature probe. Among them, NBD-ZY37 demonstrated favorable stability against photodegradation, and its temperature sensitivity in [...] Read more.
This study aimed to develop a two-color pressure-sensitive paint (PSP) that has both high pressure sensitivity and high temperature sensitivity. Different nitrobenzoxadiazole (NBD) derivatives were used as the temperature probe. Among them, NBD-ZY37 demonstrated favorable stability against photodegradation, and its temperature sensitivity in an RTV118-based two-color PSP was −1.4%/°C. Moreover, temperature sensitivity was independent of pressure in the tested temperature range. PtTFPP was used, and its pressure sensitivity was measured to be 0.5% per kPa. The two-color PSP paint underwent further examination in jet impingement experiments. The experimental results indicated that the pressure fluctuation introduced by the shock waves occurred earlier at higher impingement angles. Specifically, when the pressure ratio was 2.38, increasing the impinging angle from 15° to 30° caused the location of the pressure wave to move from s/D at 0.8 to the exit of the nozzle. Simultaneously, the shape of the maximum pressure zone changed from a fan shape to a round shape. Additionally, the jet region expanded when the pressure ratio was increased. Full article
(This article belongs to the Section Aeronautics)
Show Figures

Figure 1

Figure 1
<p>Pressure calibration setup.</p>
Full article ">Figure 2
<p>Temperature calibration setup.</p>
Full article ">Figure 3
<p>Jet impingement setup.</p>
Full article ">Figure 4
<p>Geometry of the impinging jet.</p>
Full article ">Figure 5
<p>Pressure calibration in the spectrum.</p>
Full article ">Figure 6
<p>Emission spectra of two-color PSP at different temperatures.</p>
Full article ">Figure 7
<p>Calibration results from the spectrometer calibration system. (<b>a</b>) Pressure calibration and (<b>b</b>) temperature calibration.</p>
Full article ">Figure 8
<p>(<b>a</b>) Pressure calibration of a single luminophore PSP and (<b>b</b>) temperature calibration of a single luminophore PSP from monochrome and color CCD camera for comparison.</p>
Full article ">Figure 9
<p>(<b>a</b>) Pressure calibration of a single luminophore TSP and (<b>b</b>) temperature calibration of a single luminophore TSP from monochrome and color CCD camera for comparison.</p>
Full article ">Figure 10
<p>Description of pressure distribution on the surface and pressure along the centerline at α = 20° and Φ = 3.40.</p>
Full article ">Figure 11
<p>Temperature distribution on the surface and temperature along the centerline at α = 20° and Φ = 3.40.</p>
Full article ">Figure 12
<p>Calibration curves of two-color PSP from color CCD camera. (<b>a</b>) Pressure calibration and (<b>b</b>) temperature calibration.</p>
Full article ">Figure 13
<p>Pressures along the jet centerline corrected and not corrected at α = 20° and Φ = 3.40.</p>
Full article ">Figure 14
<p>Pressure distribution on the surface at α = 15°.</p>
Full article ">Figure 15
<p>Pressure distribution on the surface at α = 20°.</p>
Full article ">Figure 16
<p>Pressure distribution on the surface at α = 30°.</p>
Full article ">Figure 17
<p>Pressure distribution along the centerline at α = 15°.</p>
Full article ">Figure 18
<p>Pressure distribution along the centerline at α = 20°.</p>
Full article ">Figure 19
<p>Pressure distribution along the centerline at α = 30°.</p>
Full article ">
12 pages, 4423 KiB  
Article
Effective Distance for Vortex Generators in High Subsonic Flows
by Ping-Han Chung, Yi-Xuan Huang, Kung-Ming Chung, Chih-Yung Huang and Sergey Isaev
Aerospace 2023, 10(4), 369; https://doi.org/10.3390/aerospace10040369 - 12 Apr 2023
Cited by 2 | Viewed by 2615
Abstract
Vortex generators (VGs) are a passive method by which to alleviate boundary layer separation (BLS). The device-induced streamwise vortices propagate downstream. There is then lift-off from the surface and the vortex decays. The effectiveness of VGs depends on their geometrical configuration, spacing, and [...] Read more.
Vortex generators (VGs) are a passive method by which to alleviate boundary layer separation (BLS). The device-induced streamwise vortices propagate downstream. There is then lift-off from the surface and the vortex decays. The effectiveness of VGs depends on their geometrical configuration, spacing, and flow characteristics. In a high-speed flow regime, the VGs must be properly positioned upstream of the BLS region. Measurements using discrete pressure taps and pressure-sensitive paint (PSP) show that there is an increase in the upstream surface pressure and the downstream favorable pressure gradient. The effective distance for a flat plate in the presence of three VG configurations is determined, as is the height of the device (conventional and micro VGs). Full article
(This article belongs to the Special Issue Flow Control and Drag Reduction)
Show Figures

Figure 1

Figure 1
<p>Flow structure of a counter-rotating VG with counter-rotating vanes.</p>
Full article ">Figure 2
<p>Test configuration.</p>
Full article ">Figure 3
<p>Normalized velocity profiles for <span class="html-italic">M</span> = 0.64 and 0.83.</p>
Full article ">Figure 4
<p>Configuration of VGs: (<b>a</b>) CRV type, (<b>b</b>) Ramp type, and (<b>c</b>) CoV type [<a href="#B16-aerospace-10-00369" class="html-bibr">16</a>].</p>
Full article ">Figure 4 Cont.
<p>Configuration of VGs: (<b>a</b>) CRV type, (<b>b</b>) Ramp type, and (<b>c</b>) CoV type [<a href="#B16-aerospace-10-00369" class="html-bibr">16</a>].</p>
Full article ">Figure 5
<p>PSP measurement system.</p>
Full article ">Figure 6
<p>The calibration curve for UniFIB PSP.</p>
Full article ">Figure 7
<p>Global surface pressure pattern for <span class="html-italic">M</span> = 0.64 and <span class="html-italic">h*</span> = 1.0.</p>
Full article ">Figure 8
<p>Global surface pressure pattern for <span class="html-italic">M</span> = 0.83 and <span class="html-italic">h*</span> = 1.0.</p>
Full article ">Figure 9
<p>The effect of CoV VG on the centerline surface pressure distribution for <span class="html-italic">M</span> = 0.83.</p>
Full article ">Figure 10
<p>The maximum surface pressure coefficient upstream of the VGs.</p>
Full article ">Figure 11
<p>Pressure distribution in the presence of CoV VGs for <span class="html-italic">h*</span> = 1.0.</p>
Full article ">Figure 12
<p>Downstream pressure gradient.</p>
Full article ">Figure 13
<p>Effective distance, <span class="html-italic">L</span>/<span class="html-italic">δ</span>.</p>
Full article ">Figure 14
<p>Effective distance, <span class="html-italic">L</span>/<span class="html-italic">h</span>.</p>
Full article ">
Back to TopTop