[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (507)

Search Parameters:
Keywords = gas-turbine engine

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
22 pages, 6944 KiB  
Article
Multi-Fidelity Modelling of the Effect of Combustor Traverse on High-Pressure Turbine Temperatures
by Mario Carta, Shahrokh Shahpar, Tiziano Ghisu and Fabio Licheri
Aerospace 2024, 11(9), 750; https://doi.org/10.3390/aerospace11090750 - 12 Sep 2024
Viewed by 324
Abstract
As turbine entry temperatures of modern jet engines continue to increase, additional thermal stresses are introduced onto the high-pressure turbine rotors, which are already burdened by substantial levels of centrifugal and gas loads. Usually, for modern turbofan engines, the temperature distribution upstream of [...] Read more.
As turbine entry temperatures of modern jet engines continue to increase, additional thermal stresses are introduced onto the high-pressure turbine rotors, which are already burdened by substantial levels of centrifugal and gas loads. Usually, for modern turbofan engines, the temperature distribution upstream of the high-pressure stator is characterized by a series of high-temperature regions, determined by the circumferential arrangement of the combustor burners. The position of these high-temperature regions, both radially and circumferentially in relation to the high-pressure stator arrangement, can have a strong impact on their subsequent migration through the high-pressure stage. Therefore, for a given amount of thermal power entering the turbine, a significant reduction in maximum rotor temperatures can be achieved by adjusting the inlet temperature distribution. This paper is aimed at mitigating the maximum surface temperatures on a high-pressure turbine rotor from a modern commercial turbofan engine by conducting a parametric analysis and optimization of the inlet temperature field. The parameters considered for this study are the circumferential position of the high-temperature spots, and the overall bias of the temperature distribution in the radial direction. High-fidelity unsteady (phase-lag) and conjugate heat transfer simulations are performed to evaluate the effects of inlet clocking and radial bias on rotor metal temperatures. The optimized inlet distribution achieved a 100 K reduction in peak high-pressure rotor temperatures and 7.5% lower peak temperatures on the high-pressure stator vanes. Furthermore, the optimized temperature distribution is also characterized by a significantly more uniform heat load allocation on the stator vanes, when compared to the baseline one. Full article
(This article belongs to the Section Aeronautics)
Show Figures

Figure 1

Figure 1
<p>Schematic of the computational domain used for the “external” (<b>A</b>) and “conjugate” (<b>B</b>) flow simulations. Film cooling strip models are represented by the dashed red lines. The meaning of the acronyms can be found in nomenclature—picture distorted and not to scale.</p>
Full article ">Figure 2
<p>Total temperature contours at the HPS inlet corresponding to the different combinations of clocking and bias for the same cases listed in <a href="#aerospace-11-00750-t001" class="html-table">Table 1</a>; the dashed white lines represent the HPS LE positions (<b>A</b>), and RTDs at the HPS main inlet for the different cases in terms of non-dimensional total temperatures T0<sub>_ND</sub> as a function of span with original distribution in black, low bias in blue, and high bias in red (<b>B</b>).</p>
Full article ">Figure 3
<p>Close-up views of the computational grid over the stator hub showing RIDN holes (<b>A</b>) and view of the mesh of the rotor blade surface of the conjugate model with the LE cooling hole rows on display (<b>B</b>)—pictures distorted.</p>
Full article ">Figure 4
<p>Convergence of the phase-lag simulation in terms of stator force error (<b>A</b>) and rotor force error (<b>B</b>) monitored during the iterative calculation process.</p>
Full article ">Figure 5
<p>External phase-lag CFD, case 0 (baseline) instantaneous HPS and HPR static temperature contours at a mid-span section at cycle phase Φ = 33° with a view over the HPS–HPR sliding plane (<b>A</b>) and External Phase-lag CFD, case 0 (baseline) instantaneous HPR static temperature contours at a mid-span section at different period phases Φ (<b>B</b>)—Pictures distorted.</p>
Full article ">Figure 6
<p>External phase-lag CFD, case 0 (baseline), near-wall gas static temperatures at three constant-span sections along the HPR blade height: 25% (blue), 50% (grey), and 75% (orange) resulting from cycle-averaged solutions, plotted as a function of the section’s curvilinear co-ordinate.</p>
Full article ">Figure 7
<p>External phase-lag CFD, non-dimensional RTDs for case 0 (no bias, black), case 1 (negative bias, blue), and case 2 (positive bias, red) downstream of HPS (<b>A</b>) and downstream of HPR (<b>B</b>) resulting from cycle-averaged solutions, plotted as a function of blade span. Refer to <a href="#aerospace-11-00750-f001" class="html-fig">Figure 1</a>A for the locations of the RTDs.</p>
Full article ">Figure 8
<p>External CFD, near-wall gas static temperatures on the HPR surface for case 0 (no bias, black), case 1 (negative bias, blue), and case 2 (positive bias, red) at three constant-span sections resulting from phase-lag cycle-averaged solutions, plotted as a function of the section’s curvilinear co-ordinate.</p>
Full article ">Figure 9
<p>Conjugate heat transfer simulations, static temperatures on the HPR metal surface for case 0 (no bias, black), case 1 (negative bias, blue) and case 2 (positive bias, red) at three constant-span sections resulting from steady-state solutions, plotted as a function of the section’s curvilinear co-ordinate.</p>
Full article ">Figure 10
<p>External CFD, near-wall gas static temperatures on the HPR surface for case 0 (no bias, black), case 1 (negative bias, blue) and case 2 (positive bias, red) at three constant-span sections resulting from steady-state solutions (dashed curves) and phase-lag cycle-averaged solutions (solid curves), plotted as a function of the section’s curvilinear co-ordinate.</p>
Full article ">Figure 11
<p>External CFD, near-wall gas static temperatures on the HPR surface at 75% span at constant bias, no clocking (cases 0, 1, 2 in black), half-pitch clocking (cases 3, 4, 5 in green) resulting from phase-lag cycle-averaged solutions, plotted as a function of the section’s curvilinear co-ordinate.</p>
Full article ">Figure 12
<p>Total temperature contours at the HPS inlet corresponding to the new optimized ECU distribution (<b>A</b>) versus the best-performing parametric one, i.e., CASE 4 (<b>B</b>)—pictures distorted.</p>
Full article ">Figure 13
<p>RTDs at the HPS main inlet for the CASE 4 and ECU cases in terms of non-dimensional total temperatures (T0_<sub>ND</sub>) as a function of span, at the HPS Inlet (<b>A</b>), HPS Exit (<b>B</b>) and HPR Exit (<b>C</b>)—CASE 4 distribution (green) and ECU distribution (purple).</p>
Full article ">Figure 14
<p>External CFD, near-wall gas static temperatures on the HPR surface for CASE 4 (green) and ECU (purple) at three constant-span sections resulting from phase-lag cycle-averaged solutions, plotted as a function of the section’s curvilinear co-ordinate.</p>
Full article ">
20 pages, 7618 KiB  
Article
A Novel CEM-Based 2-DOF PID Controller for Low-Pressure Turbine Speed Control of Marine Gas Turbine Engines
by Gun-Baek So
Processes 2024, 12(9), 1916; https://doi.org/10.3390/pr12091916 - 6 Sep 2024
Viewed by 370
Abstract
Gas turbine engines have several advantages over piston reciprocating engines, such as higher output per unit volume, reduced vibration, rapid acceleration and deceleration, high power output, and clean exhaust gases. As a result, their use for propulsion in ships has been steadily increasing. [...] Read more.
Gas turbine engines have several advantages over piston reciprocating engines, such as higher output per unit volume, reduced vibration, rapid acceleration and deceleration, high power output, and clean exhaust gases. As a result, their use for propulsion in ships has been steadily increasing. However, gas turbine engines exhibit significant parameter variations depending on the rotational speed, making the design of controllers to ensure system stability while achieving satisfactory control performance, a very challenging task. In this paper, a novel CEM-based 2-DOF PID controller design technique is proposed to ensure the stability of a gas turbine engine while improving tracking and disturbance rejection performance. The proposed controller consists of a PID controller focused on enhancing disturbance rejection performance and a set-point filter to improve tracking performance. The set-point filter is composed of gains from the controller and a single weighting factor. When tuning the gains of the controller, the maximum sensitivity is considered to maintain an appropriate balance between system stability and response performance. The key novelty of this study can be summarized in two main points. One is that the controller is designed by matching characteristic equations, and by setting the roots of the desired characteristic equation as multipoles, the gains of the PID controller can be tuned with only one adjusting variable, making the tuning of the 2-DOF controller easier. The other is that the controller parameters are tuned based on maximum sensitivity, thus taking into account the robust stability of the control system. To demonstrate the feasibility of the proposed method, simulations are conducted for four scenarios using various performance indices. Full article
Show Figures

Figure 1

Figure 1
<p>The Zumwalt-class destroyer.</p>
Full article ">Figure 2
<p>The cruise ship Queen Mary 2.</p>
Full article ">Figure 3
<p>Structure of the LM2500 GT engine.</p>
Full article ">Figure 4
<p>Open-loop block diagram of the LM2500 GT engine system.</p>
Full article ">Figure 5
<p>Block diagram of an FMU.</p>
Full article ">Figure 6
<p>The fuel flow rate with respect to the LPT speed.</p>
Full article ">Figure 7
<p>The FMV spool position with respect to the LPT speed.</p>
Full article ">Figure 8
<p>The time delay with respect to the LPT speed.</p>
Full article ">Figure 9
<p>The fuel flow rate with respect to the FMV spool position.</p>
Full article ">Figure 10
<p>Comparison of the step responses between the TOPTD GT System and FOPTD model.</p>
Full article ">Figure 11
<p>Structure of the CEM based 2-DOF PID control system.</p>
Full article ">Figure 12
<p>An overall control system for speed control of a GT engine.</p>
Full article ">Figure 13
<p>Step responses and FMV spool position for the medium-speed region.</p>
Full article ">Figure 14
<p>Nyquist plots and MS circle for the medium-speed region.</p>
Full article ">Figure 15
<p>Step responses and FMV spool position for the low-speed region.</p>
Full article ">Figure 16
<p>Nyquist plots and MS circle for the low-speed region.</p>
Full article ">Figure 17
<p>Step responses and FMV spool position for the high-speed range.</p>
Full article ">Figure 18
<p>Nyquist plots and MS circle for the high-speed region.</p>
Full article ">Figure 19
<p>Step responses and FMV position for 5% parameter uncertainty.</p>
Full article ">Figure 20
<p>Nyquist plots and MS circle for 5% parameter uncertainty.</p>
Full article ">
30 pages, 13792 KiB  
Review
Modelling and Simulation of Effusion Cooling—A Review of Recent Progress
by Hao Xia, Xiaosheng Chen and Christopher D. Ellis
Energies 2024, 17(17), 4480; https://doi.org/10.3390/en17174480 - 6 Sep 2024
Viewed by 337
Abstract
Effusion cooling is often regarded as one of the critical techniques to protect solid surfaces from exposure to extremely hot environments, such as inside a combustion chamber where temperature can well exceed the metal melting point. Designing such efficient cooling features relies on [...] Read more.
Effusion cooling is often regarded as one of the critical techniques to protect solid surfaces from exposure to extremely hot environments, such as inside a combustion chamber where temperature can well exceed the metal melting point. Designing such efficient cooling features relies on thorough understanding of the underlying flow physics for the given engineering scenarios, where physical testing may not be feasible or even possible. Inevitably, under these circumstances, modelling and numerical simulation become the primary predictive tools. This review aims to give a broad coverage of the numerical methods for effusion cooling, ranging from the empirical models (often based on first principles and conservation laws) for solving the Reynolds-Averaged Navier–Stokes (RANS) equations to higher-fidelity methods such as Large-Eddy Simulation (LES) and hybrid RANS-LES, including Detached-Eddy Simulation (DES). We also highlight the latest progress in machine learning-aided and data-driven RANS approaches, which have gained a lot of momentum recently. They, in turn, take advantage of the higher-fidelity eddy-resolving datasets performed by, for example, LES or DES. The main examples of this review are focused on the applications primarily related to internal flows of gas turbine engines. Full article
(This article belongs to the Section J: Thermal Management)
Show Figures

Figure 1

Figure 1
<p>Illustrative concepts of film cooling (<b>left</b>) and effusion cooling (<b>right</b>).</p>
Full article ">Figure 2
<p>Fluid dynamicists’ impression of a single circular coolant jet in a heated crossflow, inspired by [<a href="#B4-energies-17-04480" class="html-bibr">4</a>]. Notice the presence of horseshoe and wake vortices may not be as strong at different jet angles.</p>
Full article ">Figure 3
<p>Hierarchy of fidelity and cost of CFD methods.</p>
Full article ">Figure 4
<p>A sketch of the control volume method described by Goldstein [<a href="#B1-energies-17-04480" class="html-bibr">1</a>].</p>
Full article ">Figure 5
<p>A sketch of the coolant film development theory by Baldauf and his colleagues [<a href="#B16-energies-17-04480" class="html-bibr">16</a>].</p>
Full article ">Figure 6
<p>A sketch of the semi-analytical method proposed by LeGrives [<a href="#B17-energies-17-04480" class="html-bibr">17</a>,<a href="#B18-energies-17-04480" class="html-bibr">18</a>].</p>
Full article ">Figure 7
<p>Improved centreline ACE using the anisotropic model of Li et al. [<a href="#B39-energies-17-04480" class="html-bibr">39</a>] (AASF AAEV <span class="html-italic">k</span>-<math display="inline"><semantics> <mi>ε</mi> </semantics></math>). Adapted from Int. J. Heat Mass Trans., Vol. 91, Li et al. [<a href="#B39-energies-17-04480" class="html-bibr">39</a>], Application of algebraic anisotropic turbulence models to film cooling flows, published by Elsevier.</p>
Full article ">Figure 8
<p>A sample of cooling effectiveness results from Krawcwiw [<a href="#B50-energies-17-04480" class="html-bibr">50</a>] for varying effusion cooling hole geometries. Reprinted from Krawciw [<a href="#B50-energies-17-04480" class="html-bibr">50</a>], Optimisation techniques for combustor wall cooling, published on the Loughborough University repository under the terms of a CC BY-NC-ND 4.0 license.</p>
Full article ">Figure 9
<p>Visualisation of a multi-hole effusion cooling array from a hybrid RANS-LES study. Reprinted from Appl. Therm. Eng., Vol. 184, Chen et al. [<a href="#B7-energies-17-04480" class="html-bibr">7</a>], Study of an effusion-cooled plate with high level of upstream fluctuation, published by Elsevier; CCC license for re-use obtained.</p>
Full article ">Figure 10
<p>LES solution showing turbulent structures and mixing behaviour of coolant downstream of a single row of cooling holes with an 80-million cell mesh. This figure is an original extension of the earlier work published by Ellis and Xia [<a href="#B10-energies-17-04480" class="html-bibr">10</a>].</p>
Full article ">Figure 11
<p>Visualisation of large scale turbulent structures with <math display="inline"><semantics> <msub> <mi>λ</mi> <mn>2</mn> </msub> </semantics></math> and normalised temperature for a case with no inflow turbulence (<b>a</b>,<b>b</b>), near-wall turbulent boundary layer (<b>c</b>,<b>d</b>) and freestream turbulence (<b>e</b>,<b>f</b>). Reprinted from Int. J. Heat Mass Trans., Vol. 195, Ellis and Xia [<a href="#B10-energies-17-04480" class="html-bibr">10</a>], Impact of inflow turbulence on large-eddy simulation of film cooling flows, published by Elsevier, under the terms of the CC BY 4.0 license.</p>
Full article ">Figure 12
<p>Tensor-basis neural network (TBNN) model used by Milani et al. [<a href="#B105-energies-17-04480" class="html-bibr">105</a>] and Ellis and Xia [<a href="#B11-energies-17-04480" class="html-bibr">11</a>,<a href="#B106-energies-17-04480" class="html-bibr">106</a>] for turbulent diffusivity tensors and turbulence anisotropy, respectively. Reprinted from Phys. Fluids, 35, Ellis and Xia [<a href="#B11-energies-17-04480" class="html-bibr">11</a>], Data-driven turbulence anisotropy in film and effusion cooling flows, published by AIP Publishing, under the terms of the CC BY 4.0 license.</p>
Full article ">Figure 13
<p>Spanwise averaged ACE for the multi-hole cooling array case compared to experimental datasets [<a href="#B48-energies-17-04480" class="html-bibr">48</a>] [•]. Reprinted from Phys. Fluids, 35, Ellis and Xia [<a href="#B11-energies-17-04480" class="html-bibr">11</a>], Data-driven turbulence anisotropy in film and effusion cooling flows, published by AIP Publishing, under the terms of the CC BY 4.0 license. (<b>a</b>) Experimental dataset [•], <span class="html-italic">SST-GDH</span> [<span class="html-fig-inline" id="energies-17-04480-i001"><img alt="Energies 17 04480 i001" src="/energies/energies-17-04480/article_deploy/html/images/energies-17-04480-i001.png"/></span>] and <span class="html-italic">TBNN-SST-GDH</span> [<span class="html-fig-inline" id="energies-17-04480-i002"><img alt="Energies 17 04480 i002" src="/energies/energies-17-04480/article_deploy/html/images/energies-17-04480-i002.png"/></span>]. (<b>b</b>) Experimental dataset [•], <span class="html-italic">SST-HOG</span> [<span class="html-fig-inline" id="energies-17-04480-i003"><img alt="Energies 17 04480 i003" src="/energies/energies-17-04480/article_deploy/html/images/energies-17-04480-i003.png"/></span>] and <span class="html-italic">TBNN-SST-HOG</span> [<span class="html-fig-inline" id="energies-17-04480-i004"><img alt="Energies 17 04480 i004" src="/energies/energies-17-04480/article_deploy/html/images/energies-17-04480-i004.png"/></span>].</p>
Full article ">
17 pages, 3442 KiB  
Article
Improvement of Hydrogen-Resistant Gas Turbine Engine Blades: Single-Crystal Superalloy Manufacturing Technology
by Alexander I. Balitskii, Yulia H. Kvasnytska, Ljubomyr M. Ivaskevych, Katrine H. Kvasnytska, Olexiy A. Balitskii, Radoslaw M. Miskiewicz, Volodymyr O. Noha, Zhanna V. Parkhomchuk, Valentyn I. Veis and Jakub Maciej Dowejko
Materials 2024, 17(17), 4265; https://doi.org/10.3390/ma17174265 - 28 Aug 2024
Viewed by 437
Abstract
This paper presents the results of an analysis of resistance to hydrogen embrittlement and offers solutions and technologies for manufacturing castings of components for critical applications, such as blades for gas turbine engines (GTEs). The values of the technological parameters for directional crystallization [...] Read more.
This paper presents the results of an analysis of resistance to hydrogen embrittlement and offers solutions and technologies for manufacturing castings of components for critical applications, such as blades for gas turbine engines (GTEs). The values of the technological parameters for directional crystallization (DC) are determined, allowing the production of castings with a regular dendritic structure of the crystallization front in the range of 10 to 12 mm/min and a temperature gradient at the crystallization front in the range of 165–175 °C/cm. The technological process of making GTE blades has been improved by using a scheme for obtaining disposable models of complex profile castings with the use of 3D printing for the manufacture of ceramic molds. The ceramic mold is obtained through an environmentally friendly technology using water-based binders. Short-term tensile testing of the samples in gaseous hydrogen revealed high hydrogen resistance of the CM-88 alloy produced by directed crystallization technology: the relative elongation in hydrogen at a pressure of 30 MPa increased from 2% for the commercial alloy to 8% for the experimental single-crystal alloy. Full article
(This article belongs to the Collection Machining and Manufacturing of Alloys and Steels)
Show Figures

Figure 1

Figure 1
<p>External view of the experimental sample.</p>
Full article ">Figure 2
<p>Exterior (<b>a</b>) and scheme (<b>b</b>) of the casting setup with an additional cooling unit: 1—casting mold; 2—vacuum jacket (vacuum chamber); 3—crystallizer; 4—heating zone of the technological chamber; 5—connection pipe for the system; 6—connection pipe for connecting to the additional unloading system of the cooling zone of the working chamber; 7—molten metal; 8—crucible of the loading device with molten metal; 9—thermal insulation screen with a central hole; 10—cooling zone of the technological chamber; 11—inductors of the heating zone; 12—thermal insulation of the inductor; 13—graphite muffle furnace; 14—stopper ring of the thermal insulation screen; 15—cylindrical wall of the thermal insulation screen; 16—jacket of the technological chamber in the cooling zone; 17—ring gas collector; 18—gas ejectors; 19—holes for gas ejectors [<a href="#B34-materials-17-04265" class="html-bibr">34</a>].</p>
Full article ">Figure 2 Cont.
<p>Exterior (<b>a</b>) and scheme (<b>b</b>) of the casting setup with an additional cooling unit: 1—casting mold; 2—vacuum jacket (vacuum chamber); 3—crystallizer; 4—heating zone of the technological chamber; 5—connection pipe for the system; 6—connection pipe for connecting to the additional unloading system of the cooling zone of the working chamber; 7—molten metal; 8—crucible of the loading device with molten metal; 9—thermal insulation screen with a central hole; 10—cooling zone of the technological chamber; 11—inductors of the heating zone; 12—thermal insulation of the inductor; 13—graphite muffle furnace; 14—stopper ring of the thermal insulation screen; 15—cylindrical wall of the thermal insulation screen; 16—jacket of the technological chamber in the cooling zone; 17—ring gas collector; 18—gas ejectors; 19—holes for gas ejectors [<a href="#B34-materials-17-04265" class="html-bibr">34</a>].</p>
Full article ">Figure 3
<p>Scheme of thermocouple arrangement: 1—ceramic mold; 2—metal sample; 3—thermocouples; 4—graphite crucible; 5—ceramic substrate; mm.</p>
Full article ">Figure 4
<p>Determination of local temperature values over melting time.</p>
Full article ">Figure 5
<p>Macrostructure of the sample in the cast state: (<b>a</b>) macrostructure of the middle part of the sample; (<b>b</b>) cross-section from the seed (near the cone) [<a href="#B2-materials-17-04265" class="html-bibr">2</a>].</p>
Full article ">Figure 6
<p>Cross-sectional microstructure of a sample of heat-resistant corrosion-resistant alloy CM88.</p>
Full article ">Figure 7
<p>Microstructure of the CM-88 alloy sample at different magnifications (<b>a</b>–<b>d</b>).</p>
Full article ">
18 pages, 12211 KiB  
Article
A Study of an Integrated Analysis Model with Secondary Flow for Assessing the Performance of a Micro Turbojet Engine
by DongEun Lee, Heeyoon Chung, Young Seok Kang and Dong-Ho Rhee
Appl. Sci. 2024, 14(17), 7606; https://doi.org/10.3390/app14177606 - 28 Aug 2024
Viewed by 399
Abstract
The objective of this study is to implement a more realistic integrated analysis model for micro gas turbines by incorporating secondary flow and combustion efficiency into the existing model, which includes main engine components such as the compressor and turbine, and to validate [...] Read more.
The objective of this study is to implement a more realistic integrated analysis model for micro gas turbines by incorporating secondary flow and combustion efficiency into the existing model, which includes main engine components such as the compressor and turbine, and to validate this model by comparing it with test results. The study was based on the JetCat P300-RX, which has a maximum thrust level of 300 N. Simulations were performed using ANSYS CFX, employing the κ-ω SST turbulence model and a mixing plane interface between individual components. The eddy dissipation model (EDM), with a combustion efficiency of 90%, was used as the combustion model. A user subroutine was also applied for the power matching of the compressor and turbine to calculate the fuel flow rate in each iteration. For secondary flow, it was assumed that 3% of the total air flow rate would flow through the secondary path and be applied to the compressor and turbine. Simulations were conducted over a range of 30,000 to 104,000 RPM, with ground conditions evaluated, including altitude-simulated conditions. To validate the analysis model, engine performance metrics such as pressure ratio, air flow rate, fuel flow rate, and exhaust gas temperature (EGT) were compared with test results. The results demonstrated that errors were less than 5% for most engine performance metrics, except for EGT and fuel flow. The discrepancy in EGT was attributed to differences in the sensing methods, while the variation in fuel flow was found to be due to the lubrication system and losses due to the secondary air flow. Consequently, this study confirmed that the integrated simulation model accurately predicts engine performance. The results indicate that the integrated simulation model provides a more realistic prediction of overall engine performance compared to previous studies. Therefore, it can evaluate detailed thermo-fluid properties without the need for component performance maps, enhancing performance evaluation and analysis. Full article
(This article belongs to the Special Issue Advances and Applications of CFD (Computational Fluid Dynamics))
Show Figures

Figure 1

Figure 1
<p>P300-RX Micro Gas Turbine Engine [<a href="#B22-applsci-14-07606" class="html-bibr">22</a>].</p>
Full article ">Figure 2
<p>Micro Gas Turbine Engine Main Components [<a href="#B20-applsci-14-07606" class="html-bibr">20</a>].</p>
Full article ">Figure 3
<p>Secondary Air Flow Schematic.</p>
Full article ">Figure 4
<p>Fluid Domain of Integrated Analysis Model.</p>
Full article ">Figure 5
<p>Grid System of Integrated Analysis Model: (<b>a</b>) Compressor; (<b>b</b>) Combustor; (<b>c</b>) Turbine; (<b>d</b>) Exhaust Nozzle.</p>
Full article ">Figure 6
<p>Y+ Contour Plots at Different Engine RPM Conditions: (<b>a</b>) 50,000 RPM; (<b>b</b>)104,000 RPM.</p>
Full article ">Figure 7
<p>Grid Independence Test at 90,000 RPM Condition.</p>
Full article ">Figure 8
<p>Boundary Conditions and Interfaces of Numerical Domain.</p>
Full article ">Figure 9
<p>Impeller and Diffuser Blade Geometry.</p>
Full article ">Figure 10
<p>Secondary Air Flow Modeling: (<b>a</b>) Impeller–Diffuser; (<b>b</b>) Turbine Stator–Turbine Rotor.</p>
Full article ">Figure 11
<p>Mass flow Rate of Working Fluid for Each Components Inlet at 104,000 RPM.</p>
Full article ">Figure 12
<p>Flow Field Contour Plots Depending on Interface Conditions at 10% Span at 104,000 RPM Condition: (<b>a</b>) Frozen Rotor Interface w/o Secondary Flow; (<b>b</b>) Mixing Plane Interface w/ Secondary Flow.</p>
Full article ">Figure 13
<p>Engine Performances According to Combustion Efficiency at 100,000 RPM: (<b>a</b>) Total Pressure at Each Component; (<b>b</b>) Air and Fuel Mass Flow Rates.</p>
Full article ">Figure 14
<p>Combustor Temperature Contour Plots at 104,000 RPM: (<b>a</b>) 100% Combustion Efficiency; (<b>b</b>) 90% Combustion Efficiency.</p>
Full article ">Figure 15
<p>Streamlines of Turbine Rotor (70,000 RPM) and Static Pressure Distribution According to Secondary Air Flow: (<b>a</b>) w/o Secondary Air Flow; (<b>b</b>) w/ Secondary Air Flow; (<b>c</b>) 10% Span; (<b>d</b>) 90% Span.</p>
Full article ">Figure 16
<p>Micro Gas Turbine Ground Test Rig.</p>
Full article ">Figure 17
<p>Altitude Engine Test Facility (AETF) at KARI [<a href="#B26-applsci-14-07606" class="html-bibr">26</a>].</p>
Full article ">Figure 18
<p>Engine Performances Comparison According to Shaft Rotating Speed: (<b>a</b>) Air Flow Rate and Thrust; (<b>b</b>) Pressure Ratio and EGT.</p>
Full article ">Figure 19
<p>Engine Performances Comparison According to Shaft Rotating Speed: (<b>a</b>) Fuel Flow Rate; (<b>b</b>) TSFC.</p>
Full article ">Figure 20
<p>Temperature Profile and Streamlines around EGT Sensor Location: (<b>a</b>) 30,000 RPM; (<b>b</b>) 104,000 RPM.</p>
Full article ">Figure 21
<p>Fuel Lubrication System.</p>
Full article ">Figure 22
<p>Engine Performances Comparison Considering the Lubrication Effect According to Shaft Rotating Speed: (<b>a</b>) Fuel Flow Rate; (<b>b</b>) TSFC.</p>
Full article ">Figure 23
<p>Engine Performances According to Bleed Mass Flow Rate at 50,000 RPM: (<b>a</b>) Fuel Flow Rate; (<b>b</b>) TSFC.</p>
Full article ">Figure 24
<p>Fuel Flow Rate and Thrust at 100,000 RPM According to Altitude Conditions: (<b>a</b>) Raw Performances; (<b>b</b>) Corrected Performances [<a href="#B21-applsci-14-07606" class="html-bibr">21</a>].</p>
Full article ">
20 pages, 3309 KiB  
Article
Thermodynamic Analysis of Marine Diesel Engine Exhaust Heat-Driven Organic and Inorganic Rankine Cycle Onboard Ships
by Cuneyt Ezgi and Haydar Kepekci
Appl. Sci. 2024, 14(16), 7300; https://doi.org/10.3390/app14167300 - 19 Aug 2024
Viewed by 431
Abstract
Due to increasing emissions and global warming, in parallel with the increasing world population and energy needs, IMO has introduced severe rules for ships. Energy efficiency on ships can be achieved using the organic and inorganic Rankine cycle (RC) driven by exhaust heat [...] Read more.
Due to increasing emissions and global warming, in parallel with the increasing world population and energy needs, IMO has introduced severe rules for ships. Energy efficiency on ships can be achieved using the organic and inorganic Rankine cycle (RC) driven by exhaust heat from marine diesel engines. In this study, toluene, R600, isopentane, and n-hexane as dry fluids; R717 and R718 as wet fluids; and R123, R142b, R600a, R245fa, and R141b as isentropic fluids are selected as the working fluid because they are commonly used refrigerants, with favorable thermal properties, zero ODP, low GWP and are good contenders for this application. The cycle and exergy efficiencies, net power, and irreversibility of marine diesel engine exhaust-driven simple RC and RC with a recuperator are calculated. For dry fluids, the most efficient fluid at low turbine inlet temperatures is n-hexane at 39.75%, while at high turbine inlet temperatures, it is toluene at 41.20%. For isentropic fluids, the most efficient fluid at low turbine inlet temperatures is R123 with 23%, while at high turbine inlet temperatures it is R141b with 23%. As an inorganic fluid, R718 is one of the most suitable working fluids at high turbine inlet temperatures of 300 °C onboard ships with a safety group classification of A1, ODP of 0, and GWP100 of 0, with a cycle efficiency of 33%. This study contributes to significant improvements in fuel efficiency and reductions in greenhouse gas emissions, leading to more sustainable and cost-effective maritime operations. Full article
(This article belongs to the Special Issue Advances in Applied Marine Sciences and Engineering—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Simple RC system.</p>
Full article ">Figure 2
<p>RC system with recuperator.</p>
Full article ">Figure 3
<p>Cycle efficiency for dry fluids.</p>
Full article ">Figure 4
<p>Cycle efficiency for wet fluids.</p>
Full article ">Figure 5
<p>Cycle efficiency for isentropic fluids.</p>
Full article ">Figure 6
<p>Exergy efficiency for dry fluids.</p>
Full article ">Figure 7
<p>Exergy efficiency for wet fluids.</p>
Full article ">Figure 8
<p>Exergy efficiency for isentropic fluids.</p>
Full article ">Figure 9
<p>Net power and irreversibility for dry fluids with simple ORC.</p>
Full article ">Figure 10
<p>Net power and irreversibility for dry fluids of RC with recuperator.</p>
Full article ">Figure 11
<p>Net power and irreversibility for wet fluids.</p>
Full article ">Figure 12
<p>Net power and irreversibility for isentropic fluids.</p>
Full article ">
15 pages, 6439 KiB  
Article
Influence of High-Temperature Aggressive Environments on the Durability of Composites Reinforced with Refractory Particles
by Peter Rusinov, George Kurapov, Anastasia Rusinova, Maxim Semadeni and Polina Sereda
Metals 2024, 14(8), 939; https://doi.org/10.3390/met14080939 - 16 Aug 2024
Viewed by 585
Abstract
The problem of increasing heat resistance and corrosion and erosion resistance of gas turbine units in compressor stations was solved through the development of new layered materials containing nanostructured grains. The authors carried out a destruction analysis of gas turbine units in compressor [...] Read more.
The problem of increasing heat resistance and corrosion and erosion resistance of gas turbine units in compressor stations was solved through the development of new layered materials containing nanostructured grains. The authors carried out a destruction analysis of gas turbine units in compressor stations. It was shown that after 10–30,000 h of operation, the greatest damage occurred when the gas turbine operated in dusty environments at high temperatures (or in air environments with a high salt content). The developed layered composites include the thermal barrier and functional reinforced nanostructured layers consisting of refractory carbides and oxides. This paper describes the destruction mechanism of gas turbine units under the influence of high-temperature aggressive environments. As a result, a new formation technology for reinforced nanostructured layered composites has been developed. The developed composition makes it possible to increase the heat resistance of materials by approximately 10 times. This significantly increases the reliability and durability of gas turbine units in compressor stations. The structural and mechanical parameters of the layered nanostructured heat-resistant composites have been studied. Full article
Show Figures

Figure 1

Figure 1
<p>The general view of the gas turbine unit—(<b>a</b>); the general view of the turbine blades, after 30,000 h of operation—(<b>b</b>); the destruction scheme of gas turbine blades from exposure to high-temperature aggressive media—(<b>c</b>).</p>
Full article ">Figure 2
<p>The corrosion stages of nozzle blades (the GTD111 alloy): the second (15,000 h)—(<b>a</b>); the third (19,000 h)—(<b>b</b>); the shoulder blades’ destruction—(<b>c</b>).</p>
Full article ">Figure 3
<p>The microstructure of the GTD111 material—(<b>a</b>); the corrosion of the blade (the GTD 111 alloy)—(<b>b</b>,<b>c</b>); the corrosive destruction of the surface layer (Alloy 1)—(<b>d</b>).</p>
Full article ">Figure 4
<p>The microstructure of the layered composite: GTD111 (1)—TBL (2)—Alloy 1 (3)—(<b>a</b>); TBL (2) structure—(<b>b</b>); Alloy 1 structure—(<b>c</b>,<b>d</b>); Alloy 2 structure—(<b>e</b>,<b>f</b>); the microelectron diffraction patterns of composite layers: Alloy 1—(<b>g</b>) and Alloy 2—(<b>h</b>).</p>
Full article ">Figure 4 Cont.
<p>The microstructure of the layered composite: GTD111 (1)—TBL (2)—Alloy 1 (3)—(<b>a</b>); TBL (2) structure—(<b>b</b>); Alloy 1 structure—(<b>c</b>,<b>d</b>); Alloy 2 structure—(<b>e</b>,<b>f</b>); the microelectron diffraction patterns of composite layers: Alloy 1—(<b>g</b>) and Alloy 2—(<b>h</b>).</p>
Full article ">Figure 5
<p>The quantitative analysis of grain size in the structure of the composite materials: (<b>a</b>) Alloy 1; (<b>b</b>) Alloy 2.</p>
Full article ">Figure 6
<p>The X-ray diffraction patterns of composite materials: (<b>a</b>) TBL; (<b>b</b>) Alloy 1; (<b>c</b>) Alloy 2.</p>
Full article ">Figure 6 Cont.
<p>The X-ray diffraction patterns of composite materials: (<b>a</b>) TBL; (<b>b</b>) Alloy 1; (<b>c</b>) Alloy 2.</p>
Full article ">Figure 7
<p>The dependence of the specific change in the mass of the GTU materials on the exposure times to an aggressive environment: Inconel 738LC (1); GTD 111 (2); Rene N6 (3); Alloy 1 (4); Alloy 2 (5).</p>
Full article ">
14 pages, 3076 KiB  
Article
Design and Thermodynamic Analysis of Waste Heat-Driven Liquid Metal–Water Binary Vapor Power Plant Onboard Ship
by Haydar Kepekci and Cuneyt Ezgi
J. Mar. Sci. Eng. 2024, 12(8), 1400; https://doi.org/10.3390/jmse12081400 - 15 Aug 2024
Viewed by 431
Abstract
Day after day, stricter environmental regulations and rising operating costs and fuel prices are forcing the shipping industry to find more effective ways of designing and operating energy-efficient ships. One of the ways to produce electricity efficiently is to create a waste heat-driven [...] Read more.
Day after day, stricter environmental regulations and rising operating costs and fuel prices are forcing the shipping industry to find more effective ways of designing and operating energy-efficient ships. One of the ways to produce electricity efficiently is to create a waste heat-driven liquid metal–water binary vapor power plant. The liquid metal Rankine cycle systems could be considered topping cycles. Liquid metal binary cycles share characteristics like those of the steam Rankine power plants. They have the potential for high conversion efficiency, they will likely produce lower-cost power in plants of large capacity rather than small, and they will operate more efficiently at design capacity rather than at partial load. As a result, liquid metal topping cycles may find application primarily as base-load plants onboard ships. In this study, a waste heat-driven liquid metal–water binary vapor power plant onboard a ship is designed and thermodynamically analyzed. The waste heat onboard the vessel is the exhaust gas of the LM2500 marine gas turbine. Mercury and Cesium are selected as liquid metals in the topping cycle, while water is used in the bottoming cycle in binary power plants. Engineering Equation Solver (EES) software (V11.898) is used to perform analyses. For the turbine inlet temperature of 550 °C, while the total net work output of the binary cycle system is calculated to be 104.84 kJ/kg liquid metal and 1740.29 kJ/kg liquid metal for mercury and cesium, respectively, the efficiency of the binary cycle system is calculated to be 31.9% and 26.3% for mercury and cesium as liquid metal, respectively. This study shows that the binary cycle has a thermal efficiency of 26.32% and 31.91% for cesium and mercury, respectively, depending on liquid metal condensing pressure, and a binary cycle thermal efficiency of 25.9% and 30.9% for cesium and mercury, respectively, depending on liquid metal turbine inlet temperature, and these are possible with marine engine waste heat-driven liquid metal–water binary vapor cycles. Full article
(This article belongs to the Section Marine Energy)
Show Figures

Figure 1

Figure 1
<p>Vapor pressure as a function of temperature for working fluids.</p>
Full article ">Figure 2
<p>Liquid metal–water binary vapor power plant onboard ship.</p>
Full article ">Figure 3
<p>T-s diagram of liquid metal–water binary vapor cycle.</p>
Full article ">Figure 4
<p>Binary cycle efficiency versus liquid metal turbine inlet temperature.</p>
Full article ">Figure 5
<p>Binary cycle efficiency versus liquid metal condensing pressure.</p>
Full article ">Figure 6
<p>Total net work output versus liquid metal inlet temperature.</p>
Full article ">Figure 7
<p>Total net work output versus liquid metal condensing pressure.</p>
Full article ">Figure 8
<p>Energy supplied to the liquid metal versus liquid metal turbine inlet temperature.</p>
Full article ">Figure 9
<p>Energy supplied to the liquid metal versus liquid metal condensing pressure.</p>
Full article ">Figure 10
<p>Binary exergy efficiency versus liquid metal condensing pressure.</p>
Full article ">Figure 11
<p>Binary exergy efficiency versus liquid metal turbine inlet temperature.</p>
Full article ">
12 pages, 2324 KiB  
Article
Exploring Performance of Pyrolysis-Derived Plastic Oils in Gas Turbine Engines
by Tomasz Suchocki, Paweł Kazimierski, Katarzyna Januszewicz, Piotr Lampart, Bartosz Gawron and Tomasz Białecki
Energies 2024, 17(16), 3903; https://doi.org/10.3390/en17163903 - 7 Aug 2024
Viewed by 619
Abstract
This study explores the intersection of waste management and sustainable fuel production, focusing on the pyrolysis of plastic waste, specifically polystyrene. We examine the physicochemical parameters of the resulting waste plastic pyrolytic oils (WPPOs), blended with kerosene to form a potential alternative fuel [...] Read more.
This study explores the intersection of waste management and sustainable fuel production, focusing on the pyrolysis of plastic waste, specifically polystyrene. We examine the physicochemical parameters of the resulting waste plastic pyrolytic oils (WPPOs), blended with kerosene to form a potential alternative fuel for gas turbines. Our findings reveal that all WPPO blends lead to increased emissions, with NOX rising by an average of 61% and CO by 25%. Increasing the proportion of WPPO also resulted in a higher exhaust gas temperature, with an average rise of 12.2%. However, the thrust-specific fuel consumption (TSFC) decreased by an average of 13.8%, impacting the overall efficiency of waste-derived fuels. This study underscores the need for integrated waste-to-energy systems, bridging the gap between waste management and resource utilization. Full article
(This article belongs to the Special Issue Combustion of Alternative Fuel Blends)
Show Figures

Figure 1

Figure 1
<p>Distillation curves for the considered fuels.</p>
Full article ">Figure 2
<p>Static thrust and mass flow rate of fuel vs. turbine rotational speed for the investigated PSO/JET A blends.</p>
Full article ">Figure 3
<p>Thrust-specific fuel consumption vs. turbine rotational speed for the investigated PSO/JET A blends.</p>
Full article ">Figure 4
<p>EGT vs. turbine rotational speed for the investigated PSO/JET A blends.</p>
Full article ">Figure 5
<p>NOx<sub>T</sub> emission index vs. turbine rotational speed for the investigated PSO/JET A blends.</p>
Full article ">Figure 6
<p>CO<sub>T</sub> emission index vs. turbine rotational speed for the investigated PSO/JET A blends.</p>
Full article ">
10 pages, 5124 KiB  
Article
Thick Columnar-Structured Thermal Barrier Coatings Using the Suspension Plasma Spray Process
by Dianying Chen and Christopher Dambra
Coatings 2024, 14(8), 996; https://doi.org/10.3390/coatings14080996 - 7 Aug 2024
Cited by 1 | Viewed by 740
Abstract
Higher operating temperatures for gas turbine engines require highly durable thermal barrier coatings (TBCs) with improved insulation properties. A suspension plasma spray process (SPS) had been developed for the deposition of columnar-structured TBCs. SPS columnar TBCs are normally achieved at a short standoff [...] Read more.
Higher operating temperatures for gas turbine engines require highly durable thermal barrier coatings (TBCs) with improved insulation properties. A suspension plasma spray process (SPS) had been developed for the deposition of columnar-structured TBCs. SPS columnar TBCs are normally achieved at a short standoff distance (50.0 mm–75.0 mm), which is not practical when coating complex-shaped engine hardware since the plasma torch may collide with the components being sprayed. Therefore, it is critical to develop SPS columnar TBCs at longer standoff distances. In this work, a commercially available pressure-based suspension delivery system was used to deliver the suspension to the plasma jet, and a high-enthalpy TriplexPro-210 plasma torch was used for the SPS coating deposition. Suspension injection pressure was optimized to maximize the number of droplets injected into the hot plasma core and achieving the best particle-melting states and deposition efficiency. The highest deposition efficiency of 51% was achieved at 0.34 MPa injection pressure with a suspension flow rate of 31.0 g/min. With the optimized process parameters, 1000 μm thick columnar-structured SPS 8 wt% Y2O3-stabilized ZrO2 (8YSZ) TBCs were successfully developed at a standoff distance of 100.0 mm. The SPS TBCs have a columnar width between 100 μm and 300 μm with a porosity of ~22%. Furnace cycling tests at 1125 °C showed the SPS columnar TBCs had an average life of 1012 cycles, which is ~2.5 times that of reference air-plasma-sprayed dense vertically cracked TBCs with the same coating thickness. The superior durability of the SPS columnar TBCs can be attributed to the high-strain-tolerant microstructure. SEM cross-section characterization indicated the failure of the SPS TBCs occurred at the ceramic top coat and thermally grown oxide (TGO) interface. Full article
(This article belongs to the Special Issue Functional Coatings and Surface Science for Precision Engineering)
Show Figures

Figure 1

Figure 1
<p>Photos of LSF-400 suspension feeding system (<b>a</b>,<b>b</b>) and TriplexPro-210 suspension plasma spray (<b>c</b>).</p>
Full article ">Figure 2
<p>Effect of injection pressure on the coating deposition efficiency.</p>
Full article ">Figure 3
<p>Typical microstructures of deposits from the fixed-scan experiment: (<b>a</b>) powdery deposits at the band edge; (<b>b</b>) dark deposits at the band center; (<b>c</b>) white deposits at the intermediate band edge.</p>
Full article ">Figure 4
<p>Top views and cross-sections of SPS 8YSZ coatings at low and high magnifications. (<b>a</b>,<b>b</b>) surface morphologies; (<b>c</b>,<b>d</b>) cross-sections.</p>
Full article ">Figure 5
<p>Thermal cycle life of SPS TBCs and APS DVC TBCs.</p>
Full article ">Figure 6
<p>Photo (<b>a</b>) and SEM microstructures (<b>b</b>,<b>c</b>) of failed SPS TBCs after 1002 cycles.</p>
Full article ">Figure 7
<p>Coating coverage rate of SPS TBCs and APS DVC TBCs.</p>
Full article ">
5 pages, 174 KiB  
Editorial
Computational and Data-Driven Modeling of Combustion in Reciprocating Engines or Gas Turbines
by Maria Cristina Cameretti and Roberta De Robbio
Energies 2024, 17(16), 3863; https://doi.org/10.3390/en17163863 - 6 Aug 2024
Viewed by 583
Abstract
The targets set by the Paris Agreement to limit greenhouse gas emissions and global warming aim to significantly reduce the levels of pollutants emitted in the atmosphere from all sectors, including transportation and land use energy production [...] Full article
21 pages, 9185 KiB  
Article
Thermodynamic Analysis and Performance Evaluation of Microjet Engines in Gas Turbine Education
by Razvan Marius Catana, Grigore Cican and Gabriel-Petre Badea
Appl. Sci. 2024, 14(15), 6754; https://doi.org/10.3390/app14156754 - 2 Aug 2024
Viewed by 571
Abstract
This paper presents a detailed study on the main parameters and performance evaluation of microjet engines, at take-off regime and at various engine working regimes, based on thermodynamic analysis of a particular engine data library, from different engine manufacturers such as JetCat and [...] Read more.
This paper presents a detailed study on the main parameters and performance evaluation of microjet engines, at take-off regime and at various engine working regimes, based on thermodynamic analysis of a particular engine data library, from different engine manufacturers such as JetCat and AMT Netherlands. The studied engines have the same spool design but different thrust classes ranging from 97 to 1569 N. The particular data library includes engine specifications from catalogs or data sheets as well as our own experimental data from the JetCat P80 microjet engine, obtained using the ET 796 Jet Turbine Module, a complete testing facility for gas turbine education purposes. Various ratios and differences between certain engine main parameters and performances are studied in order to calculate values through which the analyses can be performed. Even if the engines have different thrust classes, the study examines if there are close values of the ratios and differences of parameters, that can be defined as reference parameters through which the engine performance can be compared and evaluated. Full article
(This article belongs to the Section Aerospace Science and Engineering)
Show Figures

Figure 1

Figure 1
<p>Microjet engine main stations.</p>
Full article ">Figure 2
<p>Microjet engine real cycle diagram.</p>
Full article ">Figure 3
<p>ET 796 Jet Turbine Module.</p>
Full article ">Figure 4
<p>Total specific ideal/actual ratio enthalpies for JetCat and AMT models.</p>
Full article ">Figure 5
<p>Total specific ideal/actual ratio enthalpies for JetCat P80 micro jet model.</p>
Full article ">Figure 6
<p>Actual temperatures ratio for JetCat and AMT micro jets models.</p>
Full article ">Figure 7
<p>Actual temperatures ratio for JetCat P80 micro jet model.</p>
Full article ">Figure 8
<p>Ideal temperatures ratio for JetCat and AMT micro jet models.</p>
Full article ">Figure 9
<p>Ideal temperatures ratio for JetCat P80 micro jet model.</p>
Full article ">Figure 10
<p>Percentage variation of powers for JetCat and AMT models.</p>
Full article ">Figure 11
<p>Percentage variation of powers for JetCat P80 model.</p>
Full article ">Figure 12
<p>Percentage variation of main parameters for JetCat and AMT models.</p>
Full article ">Figure 13
<p>Percentage variation of main parameters for JetCat P80 model.</p>
Full article ">Figure 14
<p>Percentage variation of jet nozzle parameters and performance for JetCat and AMT models.</p>
Full article ">Figure 15
<p>Percentage variation of jet nozzle parameters and performance for JetCat P80 model.</p>
Full article ">
32 pages, 21135 KiB  
Article
Parametric Investigation on the Influence of Turbocharger Performance Decay on the Performance and Emission Characteristics of a Marine Large Two-Stroke Dual Fuel Engine
by Haosheng Shen, Fumiao Yang, Dingyu Jiang, Daoyi Lu, Baozhu Jia, Qingjiang Liu and Xiaochi Zhang
J. Mar. Sci. Eng. 2024, 12(8), 1298; https://doi.org/10.3390/jmse12081298 - 1 Aug 2024
Viewed by 538
Abstract
Identifying and analyzing the engine performance and emission characteristics under the condition of performance decay is of significant reference value for fault diagnosis, condition-based maintenance, and health status monitoring. However, there is a lack of relevant research on the currently popular marine large [...] Read more.
Identifying and analyzing the engine performance and emission characteristics under the condition of performance decay is of significant reference value for fault diagnosis, condition-based maintenance, and health status monitoring. However, there is a lack of relevant research on the currently popular marine large two-stroke dual fuel (DF) engines. To fill the research gap, a detailed zero-/one-dimensional (0D/1D) model of a marine two-stroke DF engine employing the low-pressure gas concept is first established in GT-Power (Version 2020) and validated by comparing the simulation and measured results. Then, three typical types of turbocharger performance decays are defined including turbine efficiency decay, turbine nozzle ring area decay, and turbocharger shaft mechanical efficiency decay. Finally, the three types of decays are introduced to the engine simulation model and parametric runs are performed in both diesel and gas modes to identify and analyze their impacts on the performance and emission characteristics of the investigated marine DF engine. The results reveal that turbocharger performance decay has a significant impact on engine performance parameters, such as brake efficiency, engine speed, boost pressure, etc., as well as CO2 and NOx emissions, and the specified limit value on certain engine operational parameters will be exceeded when turbocharger performance decays to a certain extent. The changing trend of engine performance and emission parameters as turbocharger performance deteriorates are generally consistent in both operating modes but with significant differences in the extent and magnitude, mainly due to the distinct combustion process (Diesel cycle versus Otto cycle). Furthermore, considering the relative decline in brake efficiency, engine speed drop, and relative increase in CO2 emission, the investigated engine is less sensitive to the turbocharger performance decay in gas mode. The simulation results also imply that employing a variable geometry turbine (VGT) is capable of improving the brake efficiency of the investigated marine DF engine. Full article
(This article belongs to the Special Issue Performance and Emission Characteristics of Marine Engines)
Show Figures

Figure 1

Figure 1
<p>Engine simulation model in GT-Power.</p>
Full article ">Figure 2
<p>Extension of compressor performance map into the surge region.</p>
Full article ">Figure 3
<p>Comparison between simulation results and measured results in both diesel and gas modes.</p>
Full article ">Figure 4
<p>Simulation results of engine operational parameters not measured in the shop trial.</p>
Full article ">Figure 5
<p>Compressor operating point at different engine loads in both diesel and gas modes.</p>
Full article ">Figure 6
<p>In-cylinder average temperature at 75% load in both modes.</p>
Full article ">Figure 7
<p>Fraction of burned fuel at 75% load in both modes.</p>
Full article ">Figure 8
<p>Influence of turbine efficiency decay on engine performance and emission characteristics in diesel mode.</p>
Full article ">Figure 9
<p>Influence of turbine efficiency decay on engine performance and emission characteristics in gas mode.</p>
Full article ">Figure 10
<p>Influence of turbine efficiency decay on the compressor operating point.</p>
Full article ">Figure 11
<p>Influence of turbine nozzle ring area decay on engine performance and emission characteristics in diesel mode.</p>
Full article ">Figure 12
<p>Influence of turbine nozzle ring area decay on engine performance and emission characteristics in gas mode.</p>
Full article ">Figure 13
<p>Influence of turbine nozzle ring area decay on compressor operating points for both modes.</p>
Full article ">Figure 14
<p>Fitting result of the nozzle ring area multipliers and surge margins at each load condition for both modes.</p>
Full article ">Figure 15
<p>P–V diagrams with different nozzle ring area multipliers at 75% load in diesel mode.</p>
Full article ">Figure 16
<p>Variation trend in brake efficiency and turbocharger efficiency with the change in turbine nozzle ring area at 75% load for both modes.</p>
Full article ">
24 pages, 19745 KiB  
Article
Simulation and Experimental Study of Gas Turbine Blade Tenon-Root Detachment on Spin Test
by Maoyu Yu, Jianfang Wang, Haijun Xuan, Wangjiao Xiong, Zekan He and Mingmin Qu
Aerospace 2024, 11(8), 629; https://doi.org/10.3390/aerospace11080629 - 1 Aug 2024
Viewed by 556
Abstract
This paper addresses the critical issue of turbine blade containment in aircraft engines, crucial for ensuring flight safety. Through a comprehensive approach integrating numerical simulations and experimental validations, the containment capabilities of gas turbine engine casings are thoroughly analyzed. The study investigates the [...] Read more.
This paper addresses the critical issue of turbine blade containment in aircraft engines, crucial for ensuring flight safety. Through a comprehensive approach integrating numerical simulations and experimental validations, the containment capabilities of gas turbine engine casings are thoroughly analyzed. The study investigates the impact dynamics, deformation characteristics, and energy absorption mechanisms during blade detachment events, shedding light on the containment process. Based on the multi-stage nature of gas turbines, two different blade structures were designed for turbine blades. Utilizing finite element simulation and the Johnson–Cook constitutive equation, this study accurately simulated single-blade and dual-blade containment scenarios. The simulation results of the single blade indicate that the process of a gas turbine blade impacting the casing primarily consists of three stages. The second stage, where the tenon root strikes the casing, is identified as the main cause of casing damage. Meanwhile, in the dual-blade simulation, the second blade, influenced by the first blade, directly impacts the casing after fracturing, resulting in greater damage. Then, eight corresponding containment tests were conducted based on the simulation results, validating the accuracy of the simulation parameters. Experimental verification of simulation results further confirms the validity of the proposed containment curves, providing essential insights for optimizing casing design and enhancing the safety and reliability of aircraft engines. Full article
Show Figures

Figure 1

Figure 1
<p>Non-containment incident in the GECF34 engine turbine casing.</p>
Full article ">Figure 2
<p>The gas turbine blades A and B.</p>
Full article ">Figure 3
<p>Schematic diagram of the casing model.</p>
Full article ">Figure 4
<p>FEM model.</p>
Full article ">Figure 5
<p>Process of the turbine blade pieces impacting the casing A-1.</p>
Full article ">Figure 6
<p>Energy changes during the impact A-1.</p>
Full article ">Figure 7
<p>Process of the turbine blade pieces impacting the casing B-1.</p>
Full article ">Figure 8
<p>Energy changes during the impact B-1.</p>
Full article ">Figure 9
<p>Process of the turbine blade pieces impacting the casings A-2 and B-2.</p>
Full article ">Figure 10
<p>Energy changes in the containment process.</p>
Full article ">Figure 11
<p>The single-blade casing damage.</p>
Full article ">Figure 12
<p>The A-2* FEM model.</p>
Full article ">Figure 13
<p>The A-2* casing damage.</p>
Full article ">Figure 14
<p>Process of the turbine blade pieces impacting the dual-blade model casings.</p>
Full article ">Figure 14 Cont.
<p>Process of the turbine blade pieces impacting the dual-blade model casings.</p>
Full article ">Figure 15
<p>The dual-blade casing damage.</p>
Full article ">Figure 15 Cont.
<p>The dual-blade casing damage.</p>
Full article ">Figure 16
<p>Schematic diagram of the test structure.</p>
Full article ">Figure 17
<p>Containment test setup diagram.</p>
Full article ">Figure 18
<p>High-speed photography of the single-blade containment test.</p>
Full article ">Figure 18 Cont.
<p>High-speed photography of the single-blade containment test.</p>
Full article ">Figure 19
<p>Damage condition of four single-blade containment tests.</p>
Full article ">Figure 19 Cont.
<p>Damage condition of four single-blade containment tests.</p>
Full article ">Figure 20
<p>High-speed photography of the dual-blade containment test.</p>
Full article ">Figure 20 Cont.
<p>High-speed photography of the dual-blade containment test.</p>
Full article ">Figure 21
<p>Damage condition of four dual-blade containment tests.</p>
Full article ">Figure 22
<p>The comparison of the blade impact process between simulation and experiment.</p>
Full article ">Figure 23
<p>The comparison of casing damage between A-1 and B-1.</p>
Full article ">Figure 24
<p>The kinetic energy changes during the impacts A-1 and B-1.</p>
Full article ">Figure 25
<p>Turbine blade containment curve.</p>
Full article ">
21 pages, 6426 KiB  
Project Report
Fuel–Water Emulsions as an Alternative Fuel for Gas Turbines: A Project Summary
by Paweł Niszczota and Marian Gieras
Appl. Sci. 2024, 14(15), 6686; https://doi.org/10.3390/app14156686 - 31 Jul 2024
Viewed by 542
Abstract
The paper presents conclusions from research conducted at the Warsaw University of Technology in 2019–2023 regarding the combustion of fuel–water emulsions in a miniature gas turbine. The presented conclusions were made taking the current state of knowledge available in the literature into account. [...] Read more.
The paper presents conclusions from research conducted at the Warsaw University of Technology in 2019–2023 regarding the combustion of fuel–water emulsions in a miniature gas turbine. The presented conclusions were made taking the current state of knowledge available in the literature into account. Particular emphasis was placed on explaining the discrepancies in the results of the experimental studies available in the literature. The main aspects of the combustion of the fuel–water emulsions that were analyzed were their impact on the emissions of NOx and CO, as well as the impact of the surfactant included in the fuel mixture on the combustion process, emissions and the formation of deposits on the walls of the combustion chamber. The impact of the emulsion fuel on fuel consumption was also discussed. In order to explain the changes occurring in the combustion chamber as a result of adding water to the fuel, numerical methods and methods of fluid mechanics were used. Studies have shown a positive impact of the use of fuel–water emulsions on CO and NOx emissions and fuel consumption. It was also demonstrated that fuel additives used for emulsification can create deposits on the walls of the hot engine section. Full article
Show Figures

Figure 1

Figure 1
<p>Scheme of the quasi-reverse flow in a miniature gas turbine engine.</p>
Full article ">Figure 2
<p>Places of outlets of the lubrication system (GTM160) of the bearing: (<b>a</b>) front; (<b>b</b>) rear.</p>
Full article ">Figure 3
<p>Test stand: close-up of the GTM120 gas turbine.</p>
Full article ">Figure 4
<p>Cross-section of the engine with cross-section numbering (from [<a href="#B38-applsci-14-06686" class="html-bibr">38</a>]).</p>
Full article ">Figure 5
<p>Lubrication and fuel system: (<b>a</b>) unmodified; (<b>b</b>) modified.</p>
Full article ">Figure 6
<p>Scheme of the experimental trial.</p>
Full article ">Figure 7
<p>Simplified model of the GTM120 geometry.</p>
Full article ">Figure 8
<p>Photo of the fuel–water emulsion: (<b>a</b>) macroscopic; (<b>b</b>) microscopic.</p>
Full article ">Figure 9
<p>Percentage change in NOx emissions as a function of the percentage change in temperature behind the combustion chamber.</p>
Full article ">Figure 10
<p>Comparison of temperature distribution and NOx emissions at a turbine rotational speed of 80 krpm powered by standard fuel and an emulsion containing 3% water in its composition.</p>
Full article ">Figure 11
<p>The influence of water content in the emulsion on the change in CO emissions.</p>
Full article ">Figure 12
<p>Temperature distribution for the GTM120 turbine operating at a rotational speed of 100 krpm in the case of the combustion of a base fuel and an emulsion containing 12% water. (The arrow indicates the direction of the most intense expansion of the reaction area resulting from the addition of 12% water to the fuel mixture.)</p>
Full article ">Figure 13
<p>Changes in CO emissions resulting from the use of fuel–water emulsions depending on the year of publication of test results (based on data from [<a href="#B4-applsci-14-06686" class="html-bibr">4</a>]).</p>
Full article ">Figure 14
<p>The influence of water content in the emulsion on thrust-specific fuel consumption (calculated for the total mass of the fuel mixture).</p>
Full article ">Figure 15
<p>The influence of water content in the emulsion on the consumption of base fuel (after subtracting the weight of water and emulsifier).</p>
Full article ">Figure 16
<p>The influence of emulsifier content on the change in NOx emissions.</p>
Full article ">Figure 17
<p>The influence of emulsifier content on the change in CO emissions.</p>
Full article ">Figure 18
<p>GTM120 gas turbine outlet nozzle with the marked places indicating glassy deposits.</p>
Full article ">
Back to TopTop