Nothing Special   »   [go: up one dir, main page]

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

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
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
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (4,838)

Search Parameters:
Keywords = Monte-Carlo simulation

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 2713 KiB  
Article
Mineral Deposition on the Rough Walls of a Fracture
by Nathann Teixeira Rodrigues, Ismael S. S. Carrasco, Vaughan R. Voller and Fábio D. A. Aarão Reis
Minerals 2024, 14(12), 1213; https://doi.org/10.3390/min14121213 (registering DOI) - 28 Nov 2024
Abstract
Modeling carbonate growth in fractures and pores is important for understanding carbon sequestration in the environment or when supersaturated solutions are injected into rocks. Here, we study the simple but nontrivial problem of calcite growth on fractures with rough walls of the same [...] Read more.
Modeling carbonate growth in fractures and pores is important for understanding carbon sequestration in the environment or when supersaturated solutions are injected into rocks. Here, we study the simple but nontrivial problem of calcite growth on fractures with rough walls of the same mineral using kinetic Monte Carlo simulations of attachment and detachment of molecules and scaling approaches. First, we consider wedge-shaped fracture walls whose upper terraces are in the same low-energy planes and show that the valleys are slowly filled by the propagation of parallel monolayer steps in the wedge sides. The growth ceases when the walls reach these low-energy configurations so that a gap between the walls may not be filled. Second, we consider fracture walls with equally separated monolayer steps (vicinal surfaces with roughness below 1 nm) and show that growth by step propagation will eventually clog the fracture gap. In both cases, scaling approaches predict the times to attain the final configurations as a function of the initial geometry and the step-propagation velocity, which is set by the saturation index. The same reasoning applied to a random wall geometry shows that step propagation leads to lateral filling of surface valleys until the wall reaches the low-energy crystalline plane that has the smallest initial density of molecules. Thus, the final configurations of the fracture walls are much more sensitive to the crystallography than to the roughness or the local curvature. The framework developed here may be used to determine those configurations, the times to reach them, and the mass of deposited mineral. Effects of transport limitations are discussed when the fracture gap is significantly narrowed. Full article
(This article belongs to the Special Issue Mineral Dissolution and Precipitation in Geologic Porous Media)
Show Figures

Figure 1

Figure 1
<p>A region of a surface in the Kossel crystal where site colors indicate their coordinations: <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> in purple, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> in yellow, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> (kink site) in blue, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> (step site) in red, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> (terrace site) in gray, and <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math> in brown. The sites surrounding this region contain molecules that affect site colors at the boundaries.</p>
Full article ">Figure 2
<p>(<b>a</b>) Two-dimensional section of a fracture with wedge-shaped walls. The magnified zoom shows a three-dimensional view of a wedge with a small angle <math display="inline"><semantics> <mi>θ</mi> </semantics></math>, which is formed by wide terraces separated by monolayer steps (the bottoms and the tips of the wedges belong to two low-energy planes of the calcite crystal). (<b>b</b>) Two-dimensional <math display="inline"><semantics> <mrow> <mi>x</mi> <mi>z</mi> </mrow> </semantics></math> section of a fracture whose walls are vicinal surfaces forming angle <math display="inline"><semantics> <mi>θ</mi> </semantics></math> with the <span class="html-italic">z</span> direction.</p>
Full article ">Figure 3
<p>(<b>a</b>) Cross-sections (<math display="inline"><semantics> <mrow> <mi>x</mi> <mi>z</mi> </mrow> </semantics></math> plane) of a fracture with initially wedge-shaped walls, total length of 400 nm, and angle <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>15</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, in solution with <math display="inline"><semantics> <mrow> <msub> <mo>Ω</mo> <mrow> <mi>m</mi> <mi>o</mi> <mi>d</mi> <mi>e</mi> <mi>l</mi> </mrow> </msub> <mo>=</mo> <mn>3.00</mn> </mrow> </semantics></math>. In all panels, orange lines indicate the projection of the initial walls on the <math display="inline"><semantics> <mrow> <mi>x</mi> <mi>z</mi> </mrow> </semantics></math> plane. (<b>b</b>) Evolution of the bottom wall of the fracture. Site colors are those defined in <a href="#minerals-14-01213-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 4
<p>Results for growth or dissolution in wedge-shaped fracture walls: (<b>a</b>) Ratio <math display="inline"><semantics> <mrow> <mi>N</mi> <mfenced open="(" close=")"> <mi>t</mi> </mfenced> <mo>/</mo> <msub> <mi>N</mi> <mi>I</mi> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mn>400</mn> </mrow> </semantics></math> nm and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>5</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> for the saturations indicated in the plot. (<b>b</b>) Ratio <math display="inline"><semantics> <mrow> <mi>N</mi> <mfenced open="(" close=")"> <mi>t</mi> </mfenced> <mo>/</mo> <msub> <mi>N</mi> <mi>I</mi> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>5</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> for the saturations indicated in the plots. (<b>c</b>) Evolution of the growth rate for the same walls and saturations of (<b>b</b>).</p>
Full article ">Figure 5
<p>Stationary value of <math display="inline"><semantics> <mrow> <mi>N</mi> <mfenced separators="" open="(" close=")"> <msub> <mi>t</mi> <mrow> <mi>s</mi> <mi>t</mi> </mrow> </msub> </mfenced> <mo>/</mo> <msub> <mi>N</mi> <mi>I</mi> </msub> </mrow> </semantics></math> as function of <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <mi>l</mi> </mrow> </semantics></math> for different angles <math display="inline"><semantics> <mi>α</mi> </semantics></math>.</p>
Full article ">Figure 6
<p>Stationary times <math display="inline"><semantics> <msub> <mi>t</mi> <mrow> <mi>s</mi> <mi>t</mi> </mrow> </msub> </semantics></math> as function of <span class="html-italic">l</span> for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0</mn> <mo>.</mo> <msup> <mn>5</mn> <mo>∘</mo> </msup> <mo>,</mo> <mn>1</mn> <mo>.</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>5</mn> <mo>.</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Evolution of the cross-section of a fracture with vicinal surfaces with terrace length of 40 nm. The orange lines indicate the initial walls.</p>
Full article ">Figure 8
<p>Scaled time to fill the gap between the vicinal surfaces as a function of the gap distance for two different angles and saturation ratios.</p>
Full article ">Figure 9
<p>Expected time evolution of a fracture: (<b>a</b>) initial configuration with rough walls; (<b>b</b>) a configuration during calcite growth; (<b>c</b>) final configuration. Low-energy planes are indicated by parallel lines, with increasing initial density of molecules in the following order in the lower crystal: red solid line; pink dashed line; magenta dashed line; orange dashed line. Brown dashed lines in the intermediate configuration are drawn through the terraces formed around local surface peaks. Flow lines in the fracture spacing are schematically represented.</p>
Full article ">Figure 10
<p>Evolution of the cross-section of a rough wall during calcite growth. The initial configuration of the wall is indicated by the orange line and the blue line indicates the low energy plane with the smallest density of molecules in the initial wall.</p>
Full article ">Figure A1
<p>Snapshots of a growing surface (from left to right) with initially separated monolayer steps (<math display="inline"><semantics> <mrow> <mi>w</mi> <mo>/</mo> <mi>a</mi> <mo>=</mo> <mn>256</mn> </mrow> </semantics></math>) for <math display="inline"><semantics> <mrow> <msub> <mo>Ω</mo> <mrow> <mi>m</mi> <mi>o</mi> <mi>d</mi> <mi>e</mi> <mi>l</mi> </mrow> </msub> <mo>=</mo> <mn>3.00</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>=</mo> <mn>0.62</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>5.2</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math>. Site colors are the same defined in <a href="#minerals-14-01213-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure A2
<p>Monolayer step velocity as a function of the saturation ratio obtained in simulations (black circles) and AFM studies (blue filled squares [<a href="#B41-minerals-14-01213" class="html-bibr">41</a>] and red filled squares [<a href="#B42-minerals-14-01213" class="html-bibr">42</a>]).</p>
Full article ">
9 pages, 4121 KiB  
Communication
Correction of Multispectral Singlet Oxygen Luminescent Dosimetry (MSOLD) for Tissue Optical Properties in Photofrin-Mediated Photodynamic Therapy
by Weibing Yang, Madelyn Johnson, Baozhu Lu, Dennis Sourvanos, Hongjing Sun, Andreea Dimofte, Vikas Vikas, Theresa M. Busch, Robert H. Hadfield, Brian C. Wilson and Timothy C. Zhu
Antioxidants 2024, 13(12), 1458; https://doi.org/10.3390/antiox13121458 (registering DOI) - 28 Nov 2024
Abstract
The direct detection of singlet-state oxygen (1O2) constitutes the holy grail dosimetric method for type-II photodynamic therapy (PDT), a goal that can be quantified using multispectral singlet oxygen near-infrared luminescence dosimetry (MSOLD). The optical properties of tissues, specifically their [...] Read more.
The direct detection of singlet-state oxygen (1O2) constitutes the holy grail dosimetric method for type-II photodynamic therapy (PDT), a goal that can be quantified using multispectral singlet oxygen near-infrared luminescence dosimetry (MSOLD). The optical properties of tissues, specifically their scattering and absorption coefficients, play a crucial role in determining how the treatment and luminescence light are attenuated. Variations in these properties can significantly impact the spatial distribution of the treatment light and hence the generation of singlet oxygen and the detection of singlet oxygen luminescence signals. In this study, we investigated the impact of varying optical properties on the detection of 1O2 luminescence signals during Photofrin-mediated PDT in tissue-mimicking phantoms. For comparison, we also conducted Monte Carlo (MC) simulations under the same conditions. The experimental and simulations are substantially equivalent. This study advances the understanding of MSOLD during PDT. Full article
(This article belongs to the Section ROS, RNS and RSS)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>). Schematic illustration of the experimental set-up in which the laser strikes the cuvette along the z-axis and the detection fibers with NA = 0.5 are placed at 20 ± 2° to the laser beam with the fiber tip 4 ± 0.5 mm above the surface, the pink color represents the Photofrin solution, while the black color represents the black phantom. (<b>b</b>) Schematic illustration of the Monte Carlo simulation configuration, mirroring the experimental set-up, the red region represents the liquid phantom, while the blue region represents air.</p>
Full article ">Figure 2
<p>The extinction coefficient spectrum of Photofrin around the excitation (<b>a</b>) and <sup>1</sup>O<sub>2</sub> luminescence emission (<b>b</b>) wavelengths. The former is consistent with the values in the literature [<a href="#B25-antioxidants-13-01458" class="html-bibr">25</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) The absorption spectrum of the black ink from 600 nm to 1000 nm. (<b>b</b>) The calculated absorption coefficient at 632 nm as a function of the concentration, together with a linear fit with the measurements. (<b>c</b>,<b>d</b>) Corresponding measurements from 1200 to 1350 nm and at 1270 nm.</p>
Full article ">Figure 4
<p>(<b>a</b>) An example of the measured light fluence rate per unit source strength (Φ/S) at 661 nm versus the distance along the catheter from a point source: fitted <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>s</mi> </mrow> <mrow> <mo>′</mo> </mrow> </msubsup> <mo>=</mo> <mn>8.5</mn> <mtext> </mtext> <msup> <mrow> <mi mathvariant="normal">c</mi> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>. (<b>b</b>) The reduced scattering coefficient from (<b>a</b>) at 661 nm for Intralipid as a function of the concentration and the corresponding linear fit.</p>
Full article ">Figure 5
<p>Singlet oxygen luminescence spectrum of Photofrin, measured in a turbid phantom with <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo>=</mo> <mn>1.0</mn> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">c</mi> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>s</mi> </mrow> <mrow> <mo>′</mo> </mrow> </msubsup> <mo>=</mo> <mn>15</mn> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">c</mi> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>, together with the calculated component spectra.</p>
Full article ">Figure 6
<p>Measured singlet oxygen luminescence spectra with 10-point smoothing and the corresponding spectra obtained by SVD fitting. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> ranging from 0.6 to 1.5 cm<sup>−1</sup> at 632 nm (ink + Photofrin contributions) with <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>s</mi> </mrow> <mo>′</mo> </msubsup> </mrow> </semantics></math> fixed at 15 cm<sup>−1</sup>. (<b>b</b>) Corresponding spectra with <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>s</mi> </mrow> <mo>′</mo> </msubsup> </mrow> </semantics></math> = 5–40 cm<sup>−1</sup> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> = 0.5 cm<sup>−1</sup> at 632 nm.</p>
Full article ">Figure 7
<p>Dependence of the normalized <sup>1</sup>O<sub>2</sub> signal on the scattering coefficient at 632 nm, comparing the experimental data (points) and the Monte Carlo simulations (solid lines), normalized to the values of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> = 0.8 cm<sup>−1</sup> and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>s</mi> </mrow> <mrow> <mo>′</mo> </mrow> </msubsup> </mrow> </semantics></math> = 15 cm<sup>−1</sup>.</p>
Full article ">
17 pages, 2536 KiB  
Article
Analyzing Chemical Decay in Environmental Nanomaterials Using Gamma Distribution with Hybrid Censoring Scheme
by Hanan Haj Ahmad, Dina A. Ramadan and Mohamed Aboshady
Mathematics 2024, 12(23), 3737; https://doi.org/10.3390/math12233737 - 27 Nov 2024
Abstract
This study addresses the challenges of estimating decay times for chemical components, focusing on hydroxylated fullerene C60(OH)29, which poses potential environmental risks due to its persistence and transformation in soil. Given the complexities of real-world experiments [...] Read more.
This study addresses the challenges of estimating decay times for chemical components, focusing on hydroxylated fullerene C60(OH)29, which poses potential environmental risks due to its persistence and transformation in soil. Given the complexities of real-world experiments such as limited sample availability, time constraints, and the need for efficient resource use, a framework using the Gamma distribution based on hybrid Type-II censoring schemes was developed to model the decay time. The Gamma distribution’s flexibility and mathematical properties make it well-suited for reliability and decay analysis, capturing variable hazard rates and accommodating different censoring structures. We employ maximum likelihood estimation (MLE) and Bayesian methods to estimate the model’s parameters, consequently estimating the reliability and hazard functions. The large sample theory for MLE is used to approximate variances for constructing asymptotic confidence intervals. Additionally, we utilize the Markov chain Monte Carlo technique within the Bayesian framework to ensure robust parameter estimation. Through simulation studies and statistical tests—such as Chi-Square, Kolmogorov–Smirnov, and others—we assess the Gamma distribution’s fit and compare its performance with other distributions, validating the proposed model’s effectiveness. Full article
15 pages, 447 KiB  
Article
Tensor-Based Predictor–Corrector Algorithm for Power Generation and Transmission Reliability Assessment with Sequential Monte Carlo Simulation
by Erika Pequeno dos Santos, Beatriz Silveira Buss, Mauro Augusto da Rosa and Diego Issicaba
Energies 2024, 17(23), 5967; https://doi.org/10.3390/en17235967 - 27 Nov 2024
Abstract
The reliability assesment of large power systems, particularly when considering both generation and transmission facilities, is a computationally demanding and complex problem. The sequential Monte Carlo simulation is arguably the most versatile approach for tackling this problem. However, assessing sampled states in the [...] Read more.
The reliability assesment of large power systems, particularly when considering both generation and transmission facilities, is a computationally demanding and complex problem. The sequential Monte Carlo simulation is arguably the most versatile approach for tackling this problem. However, assessing sampled states in the sequential Monte Carlo simulation is time-intensive, rendering its use less appealing, particularly if nonlinear network representation must be deployed. In this context, this paper introduces a tensor-based predictor–corrector approach to reduce the burden of state evaluations in power generation and transmission reliability assessments. The approach allows for searching for sequences of operation points which can be assigned as success states within the sequential Monte Carlo simulation. If required, failure states are evaluated using a cross-entropy optimization algorithm designed to minimize load curtailments taking into account discrete variables. Numerical results emphasize the applicability of the developed algorithms using a small test system and the IEEE-RTS79 test system. Full article
(This article belongs to the Special Issue Energy, Electrical and Power Engineering: 3rd Edition)
Show Figures

Figure 1

Figure 1
<p>Core stages of the sequential Monte Carlo simulation.</p>
Full article ">Figure 2
<p>Proposed state evaluation.</p>
Full article ">Figure 3
<p>Two-bus test system.</p>
Full article ">Figure 4
<p>Results of the application of the NLMCS<math display="inline"><semantics> <mi>τ</mi> </semantics></math> and NLMCS<math display="inline"><semantics> <mover accent="true"> <mi>ρ</mi> <mo stretchy="false">^</mo> </mover> </semantics></math> methods for the two-bus system.</p>
Full article ">Figure 5
<p>IEEE-RTS79 test system.</p>
Full article ">Figure 6
<p>Results of the application of the NLMCS<math display="inline"><semantics> <mi>τ</mi> </semantics></math> and NLMCS<math display="inline"><semantics> <mover accent="true"> <mi>ρ</mi> <mo stretchy="false">^</mo> </mover> </semantics></math> methods for the IEEE-RTS79 system.</p>
Full article ">
31 pages, 10502 KiB  
Article
Flexible Simulation Platform for Generating Realistic Waveforms with Voltage Notches
by Joaquín E. Caicedo, Olga Zyabkina, Edwin Rivas and Jan Meyer
Appl. Sci. 2024, 14(23), 11031; https://doi.org/10.3390/app142311031 - 27 Nov 2024
Abstract
Voltage notches are steady-state sub-cycle waveform distortions caused by the normal operation of line-commutated power converters, significantly impacting power quality in industrial low-voltage (LV) networks. Despite their common occurrence, research on this phenomenon is still incipient, and realistic simulation platforms are lacking. This [...] Read more.
Voltage notches are steady-state sub-cycle waveform distortions caused by the normal operation of line-commutated power converters, significantly impacting power quality in industrial low-voltage (LV) networks. Despite their common occurrence, research on this phenomenon is still incipient, and realistic simulation platforms are lacking. This paper introduces a detailed MATLAB (R2024a)/Simulink-based simulation platform that models a benchmark low-voltage industrial installation, including a six-pulse controlled rectifier, linear loads, and a capacitor bank for power factor correction. Systematic simulations are performed with the platform to examine the sensitivity of notch characteristics to key parameters within plausible ranges, such as short-circuit power at the point of common coupling, commutation reactance, firing angle, snubber circuits, and rated power of the rectifier. In addition, parameters such as the rated power of linear loads and the compensation power of the capacitor bank are examined. Other influencing parameters including background voltage unbalance and distortion are also modeled and considered. A comparative analysis with field measurements from German industrial LV networks validates the plausibility and suitability of the simulations. Building upon this platform, a Monte Carlo simulation approach is adopted to generate extensive datasets of realistic voltage notch waveforms by randomly varying these key parameters. A case study conducted under conditions typical of German LV networks demonstrates the applicability of the simulations. To support further research, the simulation platform and exemplary synthetic waveforms are provided alongside the paper, serving as a valuable tool for testing and designing strategies for analysis, detection, and monitoring of voltage notches. Full article
(This article belongs to the Special Issue Analysis, Modelling and Simulation in Electrical Power Systems)
Show Figures

Figure 1

Figure 1
<p>Characteristics of voltage notches. (<b>a</b>) Depth, width, and area of a classical notch (adapted from ref. [<a href="#B5-applsci-14-11031" class="html-bibr">5</a>]). (<b>b</b>) Notch with commutation oscillations due to a capacitor bank with no detuning. (<b>c</b>) Notch with commutation oscillations due to snubber capacitance.</p>
Full article ">Figure 2
<p>Waveforms of a six-pulse controlled rectifier under different commutation conditions. The graphs display (from top to bottom): three-phase line currents, phase-to-neutral voltages, phase-to-phase voltages, and DC output voltage. (<b>a</b>) Ideal conditions: Commutation occurs smoothly without voltage notches. (<b>b</b>) Non-ideal conditions: Commutation leads to voltage notches in the waveforms.</p>
Full article ">Figure 3
<p>Benchmark LV installation implemented in Simulink for voltage notch simulation.</p>
Full article ">Figure 4
<p>General model of a six-pulse controlled rectifier.</p>
Full article ">Figure 5
<p>Overvoltage ratio <span class="html-italic">U<sub>rm</sub></span>/<span class="html-italic">U<sub>r</sub></span> as a function of the normalized snubber resistor <span class="html-italic">R<sub>sn</sub></span>/<span class="html-italic">R<sub>base</sub></span> for RC-snubbers. Curves for thyristors used in LV applications (adapted from ref. [<a href="#B21-applsci-14-11031" class="html-bibr">21</a>]).</p>
Full article ">Figure 6
<p><span class="html-italic">I<sub>rm</sub></span> versus <span class="html-italic">−di</span>/<span class="html-italic">dt</span> for thyristors in LV applications (adapted from ref. [<a href="#B23-applsci-14-11031" class="html-bibr">23</a>]).</p>
Full article ">Figure 7
<p>Conceptual diagram illustrating the interrelationships among key parameters.</p>
Full article ">Figure 8
<p>Impact of (<b>a</b>) <span class="html-italic">S<sub>sc</sub></span>, (<b>b</b>) <span class="html-italic">S<sub>r</sub></span>, (<b>c</b>) <span class="html-italic">X<sub>com</sub></span>, and (<b>d</b>) <span class="html-italic">α</span> on voltage notch waveforms (<b>top</b>) and input line currents (<b>bottom</b>).</p>
Full article ">Figure 8 Cont.
<p>Impact of (<b>a</b>) <span class="html-italic">S<sub>sc</sub></span>, (<b>b</b>) <span class="html-italic">S<sub>r</sub></span>, (<b>c</b>) <span class="html-italic">X<sub>com</sub></span>, and (<b>d</b>) <span class="html-italic">α</span> on voltage notch waveforms (<b>top</b>) and input line currents (<b>bottom</b>).</p>
Full article ">Figure 9
<p>Impact of (<b>a</b>) RC-snubber and (<b>b</b>) <span class="html-italic">R<sub>sn</sub></span> on voltage notch waveforms (<b>top</b>) and input line currents (<b>bottom</b>).</p>
Full article ">Figure 10
<p>Impact of <span class="html-italic">S<sub>ll</sub></span> = <span class="html-italic">P<sub>ll</sub></span> + <span class="html-italic">jQ<sub>ll</sub></span>, with a fixed <span class="html-italic">PF<sub>ll</sub></span> = 0.9, on voltage notch waveforms (<b>top</b>), input line currents of the rectifier (<b>middle</b>), and total line currents of the LV installation (<b>bottom</b>).</p>
Full article ">Figure 11
<p>Impact of the capacitor bank <span class="html-italic">Q<sub>cb</sub></span> on voltage notch waveforms (<b>top</b>), input line currents of the rectifier (<b>middle</b>), and total line currents of the LV installation (<b>bottom</b>). (<b>a</b>) Non-detuned capacitor bank and (<b>b</b>) detuned capacitor bank.</p>
Full article ">Figure 12
<p>Impact of α, with linear loads enabled, while maintaining a target <span class="html-italic">PF</span> at the PCC by adjusting <span class="html-italic">Q<sub>cb</sub></span> on voltage notch waveforms (<b>top</b>), input line currents of the rectifier (<b>middle</b>), and total line currents (<b>bottom</b>). (<b>a</b>) Non-detuned capacitor bank and (<b>b</b>) detuned capacitor bank.</p>
Full article ">Figure 13
<p>Impact of background unbalance on voltage notch waveforms (<b>top</b>), input line currents of the rectifier (<b>middle</b>), and total line currents of the LV installation (<b>bottom</b>). (<b>a</b>) Non-detuned capacitor bank and (<b>b</b>) detuned capacitor bank.</p>
Full article ">Figure 14
<p>Impact of background distortion on voltage notch waveforms (<b>top</b>), input line currents of the rectifier (<b>middle</b>), and total line currents of the LV installation (<b>bottom</b>). (<b>a</b>) Non-detuned capacitor bank and (<b>b</b>) detuned capacitor bank.</p>
Full article ">Figure 15
<p>Case study from field measurements.</p>
Full article ">Figure 16
<p>Example of three-phase voltage (<b>top</b>) and current (<b>bottom</b>) waveforms from a field measurement record (<b>left</b>) and a simulation (<b>right</b>).</p>
Full article ">Figure 17
<p>Overlaid voltage notch (<b>top</b>) and current (<b>bottom</b>) waveforms from the example of the field measurement record and simulation in <a href="#applsci-14-11031-f016" class="html-fig">Figure 16</a> for visual comparison. Phase <span class="html-italic">a</span> (<b>left</b>), phase <span class="html-italic">b</span> (<b>middle</b>), and phase <span class="html-italic">c</span> (<b>right</b>).</p>
Full article ">Figure 18
<p>Histograms of features extracted from field measurements.</p>
Full article ">Figure 19
<p>Matrix of scatter plots of features from field measurements and simulations.</p>
Full article ">Figure 20
<p>Correlation matrices of features extracted from field measurements (<b>left</b>) and simulations (<b>middle</b>) and normalized absolute error matrix (<b>right</b>).</p>
Full article ">Figure 21
<p>Flow diagram of Monte Carlo simulation-based data generation.</p>
Full article ">Figure 22
<p>Examples of voltage notch (<b>top</b>) and current waveforms drawn by the LV installation (<b>bottom</b>) obtained from the Monte Carlo simulation for the case study. (<b>a</b>) Scenario 1, (<b>b</b>) Scenario 2, (<b>c</b>) Scenario 3 with non-detuned capacitor bank, (<b>d</b>) Scenario 3 with detuned capacitor bank, (<b>e</b>) Scenario 4 with non-detuned capacitor bank, and (<b>f</b>) Scenario 4 with detuned capacitor bank.</p>
Full article ">
19 pages, 2858 KiB  
Article
A Resolution-Improving Method for Multiband Imaging Based on an Extrapolated RELAX Algorithm
by Jiajie Huang, Wen Jiang, Jianwei Liu, Qinyu Xie and Wangzhe Li
Remote Sens. 2024, 16(23), 4446; https://doi.org/10.3390/rs16234446 - 27 Nov 2024
Abstract
A resolution-improving method for multiband imaging based on an extrapolated RELAX algorithm (ERA) is proposed. The proposed method improves image resolution by reconstructing a wideband signal through nonadjacent narrow subbands. The key points of this method are the coherent compensation of the subbands [...] Read more.
A resolution-improving method for multiband imaging based on an extrapolated RELAX algorithm (ERA) is proposed. The proposed method improves image resolution by reconstructing a wideband signal through nonadjacent narrow subbands. The key points of this method are the coherent compensation of the subbands and the accurate extraction of target point scatters’ parameters from each subband. For coherent compensation, a two-step phase compensation (TSPC) is used to precisely compensate for the phase incoherence terms between the subbands. In order to accurately extract the parameters of point scatters (PSs), the ERA is proposed, which establishes a cost function between the extrapolated subbands and the PSs’ parameters. Furthermore, to enhance the robustness of PSs’ parameters extraction, a local band purification (LBP) process is proposed to resist the non-Gaussian clutter in the subbands. To validate the performance of the proposed method, Monte Carlo experiments using simulated data are carried out, whose results show that the proposed method can successfully improve the image resolution through the subbands under low signal-to-clutter ratio (SCR). Moreover, experiments with real measured data are conducted, using two 0.75 GHz subbands to reconstruct a 5 GHz signal, whose 2D image result is approximated to that based on the real 5 GHz signal. Image parameters are also compared, whose results demonstrate that the proposed method has satisfactory accuracy and robustness on improving image resolution. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Distribution of the subbands in frequency domain for (<b>a</b>) general condition and (<b>b</b>) simple condition.</p>
Full article ">Figure 2
<p>The major steps of the proposed method.</p>
Full article ">Figure 3
<p>Flowchart of the proposed method. Part 1: coarse phase compensation and non-Gaussian clutter decrease; part 2: PSs’ parameters estimation and fine phase compensation; part 3: FFB reconstruction.</p>
Full article ">Figure 4
<p>Range profiles from (<b>a</b>) the initial subbands, (<b>b</b>) the subbands after the LBP, (<b>c</b>) the subbands after the fine phase compensation, (<b>d</b>) the subbands and the FFB by the ERA, and (<b>e</b>) the FB and the FFB by different methods.</p>
Full article ">Figure 5
<p>RMSEs for (<b>a</b>) the proposed method, (<b>b</b>) the MRA, (<b>c</b>) the RA, and (<b>d</b>) the RMUSIC algorithm.</p>
Full article ">Figure 6
<p>(<b>a</b>) Distribution of PSs and (<b>b</b>) 2D microwave image of the FB.</p>
Full article ">Figure 7
<p>Two-dimensional microwave images based on (<b>a</b>) subband 1, (<b>b</b>) the FB, (<b>c</b>) the FFB obtained by the MRA, and (<b>d</b>) the FFB obtained by the proposed method.</p>
Full article ">Figure 8
<p>(<b>a</b>) Optical image, microwave image from (<b>b</b>) subband1, (<b>c</b>) the FFB obtained by the MRA, (<b>d</b>) the FFB obtained by the proposed method, and (<b>e</b>) the real 5 GHz FB.</p>
Full article ">
28 pages, 3873 KiB  
Article
Bayesian Inference for Long Memory Stochastic Volatility Models
by Pedro Chaim and Márcio Poletti Laurini
Econometrics 2024, 12(4), 35; https://doi.org/10.3390/econometrics12040035 - 27 Nov 2024
Abstract
We explore the application of integrated nested Laplace approximations for the Bayesian estimation of stochastic volatility models characterized by long memory. The logarithmic variance persistence in these models is represented by a Fractional Gaussian Noise process, which we approximate as a linear combination [...] Read more.
We explore the application of integrated nested Laplace approximations for the Bayesian estimation of stochastic volatility models characterized by long memory. The logarithmic variance persistence in these models is represented by a Fractional Gaussian Noise process, which we approximate as a linear combination of independent first-order autoregressive processes, lending itself to a Gaussian Markov Random Field representation. Our results from Monte Carlo experiments indicate that this approach exhibits small sample properties akin to those of Markov Chain Monte Carlo estimators. Additionally, it offers the advantages of reduced computational complexity and the mitigation of posterior convergence issues. We employ this methodology to estimate volatility dependency patterns for both the SP&500 index and major cryptocurrencies. We thoroughly assess the in-sample fit and extend our analysis to the construction of out-of-sample forecasts. Furthermore, we propose multi-factor extensions and apply this method to estimate volatility measurements from high-frequency data, underscoring its exceptional computational efficiency. Our simulation results demonstrate that the INLA methodology achieves comparable accuracy to traditional MCMC methods for estimating latent parameters and volatilities in LMSV models. The proposed model extensions show strong in-sample fit and out-of-sample forecast performance, highlighting the versatility of the INLA approach. This method is particularly advantageous in high-frequency contexts, where the computational demands of traditional posterior simulations are often prohibitive. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of stochastic volatility models’ implementation.</p>
Full article ">Figure 2
<p>Returns—Bitcoin (btc), Ethereum (eth), and S&amp;P 500 (sp500).</p>
Full article ">Figure 3
<p>Bitcoin. (btc)—Fitted volatility and VaR (5%)—AR1SV and LMSV models.</p>
Full article ">Figure 4
<p>Bitcoin (btc)—Fitted volatility and VaR (5%)—Ar1SV spline and LMSV spline models.</p>
Full article ">Figure 5
<p>S&amp;P 500 (sp500)—fitted volatility and VaR (5%)—AR1SV and LMSV models.</p>
Full article ">Figure 6
<p>S&amp;P 500 (sp500)—fitted volatility and VaR (5%)—Ar1SV spline and LMSV spline models.</p>
Full article ">Figure 7
<p>10-min intraday Bitcoin returns.</p>
Full article ">Figure 8
<p>Fitted Intraday Volatility.</p>
Full article ">Figure 9
<p>Fitted Intraday Volatility Seasonality.</p>
Full article ">
39 pages, 3120 KiB  
Article
A Comparative Review of the SWEET Simulator: Theoretical Verification Against Other Simulators
by Amine Ben-Daoued, Frédéric Bernardin and Pierre Duthon
J. Imaging 2024, 10(12), 306; https://doi.org/10.3390/jimaging10120306 - 27 Nov 2024
Viewed by 46
Abstract
Accurate luminance-based image generation is critical in physically based simulations, as even minor inaccuracies in radiative transfer calculations can introduce noise or artifacts, adversely affecting image quality. The radiative transfer simulator, SWEET, uses a backward Monte Carlo approach, and its performance is analyzed [...] Read more.
Accurate luminance-based image generation is critical in physically based simulations, as even minor inaccuracies in radiative transfer calculations can introduce noise or artifacts, adversely affecting image quality. The radiative transfer simulator, SWEET, uses a backward Monte Carlo approach, and its performance is analyzed alongside other simulators to assess how Monte Carlo-induced biases vary with parameters like optical thickness and medium anisotropy. This work details the advancements made to SWEET since the previous publication, with a specific focus on a more comprehensive comparison with other simulators such as Mitsuba. The core objective is to evaluate the precision of SWEET by comparing radiometric quantities like luminance, which serves as a method for validating the simulator. This analysis is particularly important in contexts such as automotive camera imaging, where accurate scene representation is crucial to reducing noise and ensuring the reliability of image-based systems in autonomous driving. By focusing on detailed radiometric comparisons, this study underscores SWEET’s ability to minimize noise, thus providing high-quality imaging for advanced applications. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic overview of the most important surface scattering models in SWEET.</p>
Full article ">Figure 2
<p>Comparing SWEET to Steven’s Monte Carlo code for a simple situation on fluence evaluation, which is computed as the integral of the radiance over the solid angles of 3D space <math display="inline"><semantics> <mrow> <msub> <mo>∫</mo> <mo>Ω</mo> </msub> <mi>L</mi> <mrow> <mo>(</mo> <mi>r</mi> <mo>,</mo> <mover accent="true"> <mi>u</mi> <mo>→</mo> </mover> <mo>)</mo> </mrow> <mi>d</mi> <mo>Ω</mo> </mrow> </semantics></math> (<span class="html-italic">L</span> is the radiance at position <span class="html-italic">r</span> and in a direction <math display="inline"><semantics> <mover accent="true"> <mi>u</mi> <mo>→</mo> </mover> </semantics></math>), for asymetry phase parameter: g = 0.0, optical parameters: <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 0.5 m<sup>−1</sup> and <math display="inline"><semantics> <mi>κ</mi> </semantics></math> = 0.5 m<sup>−1</sup>, the 95% confidence intervals for SWEET are in grey.</p>
Full article ">Figure 3
<p>Comparing SWEET to Mitsuba for a simple situation on radiance evaluation at a distance of 1 mm, for asymetry phase parameter: g = 0.5, optical parameters: <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 0.9 m<sup>−1</sup> and <math display="inline"><semantics> <mi>κ</mi> </semantics></math> = 0.1 m<sup>−1</sup>.</p>
Full article ">Figure 4
<p>Relative discrepancies for fluence between SWEET to Steven’s Monte Carlo code for a set of distances, optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>) and anisotropy phase parameter (g). Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </semantics></math> is taken equal to 1.0; g = 0 in the <b>left</b> and g = 0.9 in the <b>right</b>.</p>
Full article ">Figure 5
<p>Relative 95% confidence intervals of SWEET for fluence for a set of distances, optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>) and anisotropy phase parameter (g). Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, g = 0 in the <b>left</b> and g = 0.9 in the <b>right</b>.</p>
Full article ">Figure 6
<p>Comparing SWEET to Mitsuba Monte Carlo code for luminance for a set of distances and optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>). Anisotropy phase parameter (g) equals to 0.0, for the case of punctual light source. Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </semantics></math> is taken equal to 1.0, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 7
<p>Comparing SWEET to Mitsuba Monte Carlo code for luminance for a set of distances and optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>). Anisotropy phase parameter (g) equals to 0.9, for the case of punctual light source. Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </semantics></math> is taken equal to 1.0, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 8
<p>SWEET relative 95% confidence interval for luminance for a set of distances and optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>). Anisotropy phase parameter (g) equals to 0.0, for the case of punctual light source. Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </semantics></math> is taken equal to 1.0, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 9
<p>SWEET relative 95% confidence interval for luminance for a set of distances and optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>). Anisotropy phase parameter (g) equals to 0.9, for the case of punctual light source. Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </semantics></math> is taken equal to 1.0, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 10
<p>Comparing SWEET to Mitsuba Monte Carlo code for luminance for a set of distances and optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>). Anisotropy phase parameter (g) equals to 0.0, for the case of rectangular light source. Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </semantics></math> is taken equal to 1.0, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 11
<p>Comparing SWEET to Mitsuba Monte Carlo code for luminance for luminance for a set of distances and optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>). Anisotropy phase parameter (g) equals to 0.9, for the case of rectangular light source. Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </semantics></math> is taken equal to 1.0, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 12
<p>SWEET relative 95% confidence interval for a set of distances and optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>). Anisotropy phase parameter (g) equals to 0.0, for the case of rectangular light source. Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </semantics></math> is taken equal to 1.0, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 13
<p>SWEET relative 95% confidence interval for luminance for a set of distances and optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>). Anisotropy phase parameter (g) equals to 0.9, for the case of rectangular light source. Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </semantics></math> is taken equal to 1.0, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 14
<p>Comparing SWEET to Mitsuba Monte Carlo code for luminance for a set of distances and optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>). Anisotropy phase parameter (g) equals to 0.0, for the case of 2 point lights. Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </semantics></math> is taken equal to 1.0, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 15
<p>SWEET relative 95% confidence interval for luminance for a set of distances and optical coeficients (<math display="inline"><semantics> <mi>κ</mi> </semantics></math> &amp; <math display="inline"><semantics> <mi>σ</mi> </semantics></math>). Anisotropy phase parameter (g) equals to 0.0, for the case of 2 point lights. Albedo <math display="inline"><semantics> <mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>σ</mi> <mrow> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>σ</mi> <mo>+</mo> <mi>κ</mi> </mrow> </semantics></math> is taken equal to 1.0, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 16
<p>Mean path length estimated with SWEET and theorically (using IP) for cubes with side lengths ranging from 10 cm to 5 m, for asymetry phase parameter: g = 0.0, optical parameters: <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 1.0 m<sup>−1</sup> and <math display="inline"><semantics> <mi>κ</mi> </semantics></math> = 0.0 m<sup>−1</sup>.</p>
Full article ">Figure 17
<p>Mean path length estimated with SWEET and theorically (using IP) for spheres with radii ranging from 10 cm to 5 m, for asymetry phase parameter: g = 0.0, optical parameters: <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 1.0 m<sup>−1</sup> and <math display="inline"><semantics> <mi>κ</mi> </semantics></math> = 0.0 m<sup>−1</sup>.</p>
Full article ">Figure 18
<p>Relative errors of invariance property (IP) estimated with SWEET and theorically for cubes with side lengths ranging from 10 cm to 4 m, for varying asymetry phase parameter g and optical parameters (<math display="inline"><semantics> <mi>σ</mi> </semantics></math> and <math display="inline"><semantics> <mi>κ</mi> </semantics></math>), each figure corresponds to a side length as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.1</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>2.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>3.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>4.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 19
<p>Relative errors of invariance property (IP) estimated with SWEET and theorically for spheres with radii ranging from 10 cm to 4 m, for varying asymetry phase parameter g and optical parameters (<math display="inline"><semantics> <mi>σ</mi> </semantics></math> and <math display="inline"><semantics> <mi>κ</mi> </semantics></math>), each figure corresponds to a radius as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.1</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>2.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>3.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>4.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 20
<p>Execution time for SWEET and MS done with <math display="inline"><semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics></math> photons in luminance computing, for asymetry phase parameter: g = 0.9, optical parameters: <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 0.99 m<sup>−1</sup> and <math display="inline"><semantics> <mi>κ</mi> </semantics></math> = 0.01 m<sup>−1</sup>.</p>
Full article ">Figure 21
<p>Execution time variation with the photon count for SWEET and MS for a single luminance computing (<math display="inline"><semantics> <mrow> <mi>L</mi> <mo>(</mo> <mi>θ</mi> <mo>=</mo> <mn>0</mn> <mi>°</mi> <mo>)</mo> </mrow> </semantics></math>), for asymetry phase parameter: g = 0.9, optical parameters: <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 0.99 m<sup>−1</sup> and <math display="inline"><semantics> <mi>κ</mi> </semantics></math> = 0.01 m<sup>−1</sup>.</p>
Full article ">Figure 22
<p>Execution time variation for SWEET for luminance computing, done with <math display="inline"><semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics></math> photons for varying optical parameters, for asymetry phase parameter: g = 0.9, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 23
<p>Execution time variation for MS for luminance computing, done with <math display="inline"><semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics></math> photons for varying optical parameters, for asymetry phase parameter: g = 0.9, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">Figure 24
<p>Execution time ratio variation of MS to SWEET for luminance computing, done with <math display="inline"><semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics></math> photons for varying optical parameters, for asymetry phase parameter: g = 0.9, each figure corresponds to a distance as in this matrix: <math display="inline"><semantics> <mfenced open="[" close="]"> <mtable> <mtr> <mtd> <mrow> <mn>0.001</mn> </mrow> </mtd> <mtd> <mrow> <mn>0.5</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>1.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>5.0</mn> </mrow> </mtd> </mtr> <mtr> <mtd> <mrow> <mn>10.0</mn> </mrow> </mtd> <mtd> <mrow> <mn>20.0</mn> </mrow> </mtd> </mtr> </mtable> </mfenced> </semantics></math> m.</p>
Full article ">
17 pages, 2430 KiB  
Article
PyAMARES, an Open-Source Python Library for Fitting Magnetic Resonance Spectroscopy Data
by Jia Xu, Michael Vaeggemose, Rolf F. Schulte, Baolian Yang, Chu-Yu Lee, Christoffer Laustsen and Vincent A. Magnotta
Diagnostics 2024, 14(23), 2668; https://doi.org/10.3390/diagnostics14232668 - 27 Nov 2024
Viewed by 112
Abstract
Background/Objectives: Magnetic resonance spectroscopy (MRS) is a valuable tool for studying metabolic processes in vivo. While numerous quantification methods exist, the advanced method for accurate, robust, and efficient spectral fitting (AMARES) is among the most used. This study introduces pyAMARES, an open-source [...] Read more.
Background/Objectives: Magnetic resonance spectroscopy (MRS) is a valuable tool for studying metabolic processes in vivo. While numerous quantification methods exist, the advanced method for accurate, robust, and efficient spectral fitting (AMARES) is among the most used. This study introduces pyAMARES, an open-source Python implementation of AMARES, addressing the need for a flexible, user-friendly, and versatile MRS quantification tool within the Python ecosystem. Methods: PyAMARES was developed as a Python library, implementing the AMARES algorithm with additional features such as multiprocessing capabilities and customizable objective functions. The software was validated against established AMARES implementations (OXSA and jMRUI) using both simulated and in vivo MRS data. Monte Carlo simulations were conducted to assess robustness and accuracy across various signal-to-noise ratios and parameter perturbations. Results: PyAMARES utilizes spreadsheet-based prior knowledge and fitting parameter settings, enhancing flexibility and ease of use. It demonstrated comparable performance to existing software in terms of accuracy, precision, and computational efficiency. In addition to conventional AMARES fitting, pyAMARES supports fitting without prior knowledge, frequency-selective AMARES, and metabolite residual removal from mobile macromolecule (MM) spectra. Utilizing multiple CPU cores significantly enhances the performance of pyAMARES. Conclusions: PyAMARES offers a robust, flexible, and user-friendly solution for MRS quantification within the Python ecosystem. Its open-source nature, comprehensive documentation, and integration with popular data science tools enhance reproducibility and collaboration in MRS research. PyAMARES bridges the gap between traditional MRS fitting methods and modern machine learning frameworks, potentially accelerating advancements in metabolic studies and clinical applications. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of pyAMARES. The workflow starts with importing prior knowledge from spreadsheets (1a) and loading the FID signal (1b) to establish initial values and constraints for fitting (3). If the initial parameters are far from the actual values, users can optionally employ Hankel singular value decomposition (HSVD) or Levenberg–Marquardt (LM) initializers to optimize these starting values (2a). The FID signal can be processed directly or optionally filtered using MPFIR to focus on specific spectral regions (2b). The non-linear least-squares minimization (4) using either trust region reflective (TRR) or LM, with either default or user-defined objective functions (1c). The fitting process can be iterative—the output can be fine-tuned and used as initial parameters for subsequent iterations (7). The Cramér–Rao lower bound (CRLB) estimation (5) integrates information from both the fitting results and the linear relationships between parameters (2b). These relationships include constraints like fixed amplitude ratios or chemical shift differences between multiplet peaks. The final output (6) includes fitted parameters, their uncertainties (CRLB), and signal-to-noise ratios. Solid arrows indicate the main workflow, while dashed arrows and boxes represent optional processing steps.</p>
Full article ">Figure 2
<p>PyAMARES plotting outputs. The default output figure from the <span class="html-italic">plotAMARES</span> function shows the fit of (<b>A</b>) an in vivo brain <sup>31</sup>P MRS spectrum acquired at 7T [<a href="#B34-diagnostics-14-02668" class="html-bibr">34</a>], (<b>B</b>) a voxel of hyperpolarized <sup>129</sup>Xe MRSI acquired from healthy porcine lungs at 3T, and (<b>D</b>) a voxel of in vivo brain <sup>2</sup>H 3D MRSI spectra acquired at 3T. In (<b>A</b>,<b>B</b>,<b>D</b>), the top panels display the original spectrum (gray), the fitted spectrum (red), and the residual (green dash), with individual fitted components shown in the bottom panels. Panel (<b>A</b>) is shown with phase correction applied (<span class="html-italic">ifphase = True</span> for the <span class="html-italic">plotAMARES</span> function) for display purposes, while (<b>B</b>,<b>D</b>) are not phased. The prior knowledge for the fitting (<b>A</b>) is in <a href="#diagnostics-14-02668-t001" class="html-table">Table 1</a>. The fitting results for <sup>31</sup>P MRS (<b>A</b>), including metabolite concentrations and their respective Cramér–Rao lower bounds (CRLBs), are presented in (<b>C</b>), where green grows indicate reliable fits with CRLB &lt; 20% and red rows indicate less reliable fits. The fitting results of (<b>B</b>,<b>D</b>) are shown in <a href="#app1-diagnostics-14-02668" class="html-app">Figure S2</a>. Abbreviations: RBC, red blood cells; DHO, deuterated water; Glx, combined signals of glutamate and glutamine; PCr: phosphocreatine; PE: phosphoethenolamine; GPE: glycerophosphoethanolamine; GPC: glycerophosphocholine; Pi: inorganic phosphate; NAD, nicotinamide adenine dinucleotide; UDPG, uridine diphosphoglucose.</p>
Full article ">Figure 3
<p>Comparison of Monte Carlo simulated single-peak spectra fitting using OXSA and pyAMARES. (<b>A</b>) Ground truth for spectra simulation with fixed (red) and 3000 perturbed (various colors) parameters. Gaussian noise is omitted for clarity. (<b>B</b>) Relative bias of fitted amplitude compared to ground truth at different SNR levels. (<b>C</b>) Bias of fitted chemical shift compared to ground truth at different SNRs. (<b>D</b>) CRLB of fitted amplitude at each SNR, with the 20% threshold indicated by a green dashed line. In (<b>B</b>–<b>D</b>), blue and red represent pyAMARES and OXSA fitted results, respectively; solid patterns indicate results from spectra simulated with perturbed parameters, while hatched patterns show results from spectra simulated with fixed parameters.</p>
Full article ">Figure 4
<p>Comparison of Monte Carlo simulated in vivo human brain <sup>31</sup>P MRS spectra fitting at 7T using OXSA and different algorithms implemented in pyAMARES. (<b>A</b>) Ground truth for spectra simulation with slightly perturbed parameters. Gaussian noise is omitted for clarity. (<b>B</b>) Relative bias of peak amplitude quantification compared to ground truth. (<b>C</b>) CRLB of fitted amplitude for each peak, with the 20% threshold indicated by a green dashed line. (<b>D</b>) Pearson’s correlation coefficient (R) between OXSA and pyAMARES quantified amplitudes. Abbreviations: LM: Levenberg–Marquardt algorithm; TRR: trust region reflective algorithm; Init: Initializer using LM; PCr: phosphocreatine; PE: phosphoethenolamine; GPE: glycerophosphoethanolamine; GPC: glycerophosphocholine; Pi: inorganic phosphate; NAD, nicotinamide adenine dinucleotide; UDPG, uridine diphosphoglucose.</p>
Full article ">Figure 5
<p>Multiprocessing fitting of dynamic unlocalized <sup>31</sup>P MRS spectra of the tibialis anterior muscle at 3T using pyAMARES and comparison to OXSA. (<b>A</b>). Representative fitting results from pyAMARES (blue solid line) and OXSA (red dash line), with the differences between them shown as green dashed line. The metabolites of interest (PCr and Pi) are labeled. (<b>B</b>). Linear correlations between fitted amplitudes (a.u.), linewidths (Hz), and CRLBs obtained by pyAMARES and OXSA. Pearson’s R and the <span class="html-italic">p</span>-value for each dataset are shown in the plots. (<b>C</b>,<b>D</b>) Time courses of PCr (blue) and Pi (orange) amplitudes fitted by pyAMARES (<b>C</b>) and OXSA (<b>D</b>). The time points at which exercise and recovery start are indicated by dotted and dashed vertical lines, respectively. (<b>E</b>,<b>F</b>) Mono-exponential fitting of the PCr recovery kinetics using pyAMARES (<b>E</b>) and OXSA (<b>F</b>). The fitted equations are PC<sub>recover</sub> = 0.435 − 0.173 × e<sup>−time/44.171</sup>, R<sup>2</sup> = 0.914 for pyAMARES, and PC<sub>recover</sub> = 0.435 − 0.165 × e<sup>−time/42.523</sup>, R<sup>2</sup> = 0.928 for OXSA.</p>
Full article ">Figure 6
<p>Using AMARES for post-processing: Removal of metabolite residuals from a short echo time (TE) <sup>1</sup>H MR spectrum at 9.4T. (<b>A</b>) Upper panel: Fitting of residual metabolites (red) and the resulting macromolecule (MM) spectrum after subtraction of residual metabolite signals from the original spectrum (green). Lower panel: AMARES modeling of residual metabolite signals. (<b>B</b>) Comparison of metabolite-free MM spectra obtained by jMRUI (red) and pyAMARES (blue), showing identical results as confirmed by the flat difference spectrum (black).</p>
Full article ">
15 pages, 3742 KiB  
Article
Statistical Analysis of Positioning Errors in Long-Endurance Dual-Axis Rotary Modulation Inertial Navigation System
by Yaojin Hu, Hongwei Bian, Rongying Wang, Zhe Wen and Sha Hu
Electronics 2024, 13(23), 4671; https://doi.org/10.3390/electronics13234671 - 26 Nov 2024
Viewed by 246
Abstract
To address the uncertainty in the statistical distribution model of positioning error for the accuracy test of long-endurance inertial navigation systems, a probability distribution model adheres to the statistical rule of the radial positioning error of inertial navigation systems. The probability distribution density [...] Read more.
To address the uncertainty in the statistical distribution model of positioning error for the accuracy test of long-endurance inertial navigation systems, a probability distribution model adheres to the statistical rule of the radial positioning error of inertial navigation systems. The probability distribution density function (PDF), cumulative density function (CDF), and characteristic numbers (mean, standard deviation, root-mean-square) of the radial positioning error are derived based on the static-base positioning error of the long-range inertial navigation systems. Methods are provided for estimating the parameters of the probability distribution of the radial positioning errors. The theoretical derivation results demonstrate that the radial positioning error follows the Hoyt distribution. The distribution parameters and the number of features grow linearly with time, while the mean and standard deviation converge to 60% and 80% of the root-mean-square, respectively. Through a large-sample Monte Carlo simulation, the experimental results were consistent with the theoretical derivation results. These results indicate that the theoretical derivation results can be used to optimize the design of the long-endurance rotary modulation inertial navigation system’s accuracy test. Full article
(This article belongs to the Section Systems & Control Engineering)
Show Figures

Figure 1

Figure 1
<p>Conversion diagram between the OPEQ coordinate system and navigation (N) coordinate system.</p>
Full article ">Figure 2
<p>Diagram of the relationship between the OPEQ coordinate system and the N coordinate system.</p>
Full article ">Figure 3
<p>PDF of radial positioning error over time.</p>
Full article ">Figure 4
<p>PDF of radial positioning error in a typical moment.</p>
Full article ">Figure 5
<p>CDF of radial positioning error over time.</p>
Full article ">Figure 6
<p>CDF of radial positioning error in a typical moment.</p>
Full article ">Figure 7
<p>The trend of the random variable characteristic number of Hoyt distributions over time (<math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 0.001°/h).</p>
Full article ">Figure 8
<p>Comparison of PDF and Hoyt distributions of radial positioning errors at different times under static base simulation conditions.</p>
Full article ">Figure 9
<p>The statistical results of cumulative probability distribution of radial positioning error and maximum value using different calculation methods.</p>
Full article ">
17 pages, 1117 KiB  
Article
Adaptive Beamforming, Cell-Free Resource Allocation and NOMA in Large-Scale Wireless Networks
by Panagiotis Gkonis, Spyros Lavdas, George Vardoulias, Panagiotis Trakadas, Lambros Sarakis and Konstantinos Papadopoulos
Sensors 2024, 24(23), 7548; https://doi.org/10.3390/s24237548 - 26 Nov 2024
Viewed by 193
Abstract
The goal of the study presented in this work is to evaluate the performance of a proposed adaptive beamforming approach when combined with non-orthogonal multiple access (NOMA) in cell-free massive multiple input multiple output (CF m-MIMO) orientations. In this context, cooperative beamforming is [...] Read more.
The goal of the study presented in this work is to evaluate the performance of a proposed adaptive beamforming approach when combined with non-orthogonal multiple access (NOMA) in cell-free massive multiple input multiple output (CF m-MIMO) orientations. In this context, cooperative beamforming is employed taking into consideration the geographically adjacent access points (APs) of a virtual cell, aiming to minimize co-channel interference (CCI) among mobile stations (MSs) participating in NOMA transmission. Performance is evaluated statistically via extensive Monte Carlo (MC) simulations in a two-tier wireless orientation. As the results indicate, for high data rate services, various key performance indicators (KPIs) can be improved compared to orthogonal multiple access, such as the minimum number of users in the topology as well as the available PRBs for downlink transmission. Although in NOMA transmission more directional beamforming configurations are required to compensate for the increased CCI levels, the increase in the number of hardware elements is reduced compared to the corresponding gain in the considered KPIs. Full article
(This article belongs to the Special Issue Intelligent Massive-MIMO Systems and Wireless Communications)
17 pages, 377 KiB  
Article
MODE: Minimax Optimal Deterministic Experiments for Causal Inference in the Presence of Covariates
by Shaohua Xu, Songnan Liu and Yongdao Zhou
Entropy 2024, 26(12), 1023; https://doi.org/10.3390/e26121023 - 26 Nov 2024
Viewed by 152
Abstract
Data-driven decision-making has become crucial across various domains. Randomization and re-randomization are standard techniques employed in controlled experiments to estimate causal effects in the presence of numerous pre-treatment covariates. This paper quantifies the worst-case mean squared error of the difference-in-means estimator as a [...] Read more.
Data-driven decision-making has become crucial across various domains. Randomization and re-randomization are standard techniques employed in controlled experiments to estimate causal effects in the presence of numerous pre-treatment covariates. This paper quantifies the worst-case mean squared error of the difference-in-means estimator as a generalized discrepancy of covariates between treatment and control groups. We demonstrate that existing randomized or re-randomized experiments utilizing Monte Carlo methods are sub-optimal in minimizing this generalized discrepancy. To address this limitation, we introduce a novel optimal deterministic experiment based on quasi-Monte Carlo techniques, which effectively minimizes the generalized discrepancy in a model-independent manner. We provide a theoretical proof indicating that the difference-in-means estimator derived from the proposed experiment converges more rapidly than those obtained from completely randomized or re-randomized experiments using Mahalanobis distance. Simulation results illustrate that the proposed experiment significantly reduces covariate imbalances and estimation uncertainties when compared to existing randomized and deterministic approaches. In summary, the proposed experiment serves as a reliable and effective framework for controlled experimentation in causal inference. Full article
(This article belongs to the Section Information Theory, Probability and Statistics)
27 pages, 699 KiB  
Article
Estimating the Lifetime Parameters of the Odd-Generalized-Exponential–Inverse-Weibull Distribution Using Progressive First-Failure Censoring: A Methodology with an Application
by Mahmoud M. Ramadan, Rashad M. EL-Sagheer and Amel Abd-El-Monem
Axioms 2024, 13(12), 822; https://doi.org/10.3390/axioms13120822 - 25 Nov 2024
Viewed by 245
Abstract
This paper investigates statistical methods for estimating unknown lifetime parameters using a progressive first-failure censoring dataset. The failure mode’s lifetime distribution is modeled by the odd-generalized-exponential–inverse-Weibull distribution. Maximum-likelihood estimators for the model parameters, including the survival, hazard, and inverse hazard rate functions, are [...] Read more.
This paper investigates statistical methods for estimating unknown lifetime parameters using a progressive first-failure censoring dataset. The failure mode’s lifetime distribution is modeled by the odd-generalized-exponential–inverse-Weibull distribution. Maximum-likelihood estimators for the model parameters, including the survival, hazard, and inverse hazard rate functions, are obtained, though they lack closed-form expressions. The Newton–Raphson method is used to compute these estimations. Confidence intervals for the parameters are approximated via the normal distribution of the maximum-likelihood estimation. The Fisher information matrix is derived using the missing information principle, and the delta method is applied to approximate the confidence intervals for the survival, hazard rate, and inverse hazard rate functions. Bayes estimators were calculated with the squared error, linear exponential, and general entropy loss functions, utilizing independent gamma distributions for informative priors. Markov-chain Monte Carlo sampling provides the highest-posterior-density credible intervals and Bayesian point estimates for the parameters and reliability characteristics. This study evaluates these methods through Monte Carlo simulations, comparing Bayes and maximum-likelihood estimates based on mean squared errors for point estimates, average interval widths, and coverage probabilities for interval estimators. A real dataset is also analyzed to illustrate the proposed methods. Full article
Show Figures

Figure 1

Figure 1
<p>PDF for OGE-IWD.</p>
Full article ">Figure 2
<p>HRF for OGE-IWD.</p>
Full article ">Figure 3
<p>Description of the PFFC scheme.</p>
Full article ">Figure 4
<p>The KD, box, TTT, Q-Q, P-P, SF, PDF, and violin plots for the data set.</p>
Full article ">
16 pages, 2661 KiB  
Article
Pressure Transient Analysis for Fractured Shale Gas Wells Using Trilinear Flow Model
by Li Liu, Liang Xue and Jiangxia Han
Processes 2024, 12(12), 2652; https://doi.org/10.3390/pr12122652 - 25 Nov 2024
Viewed by 237
Abstract
Shale gas, a low-permeability, unconventional resource, requires horizontal drilling and multi-stage fracturing for commercial production. This study develops a trilinear flow model for fractured horizontal wells in shale gas formations, incorporating key mechanisms such as adsorption, desorption, diffusion, wellbore storage, and skin effects. [...] Read more.
Shale gas, a low-permeability, unconventional resource, requires horizontal drilling and multi-stage fracturing for commercial production. This study develops a trilinear flow model for fractured horizontal wells in shale gas formations, incorporating key mechanisms such as adsorption, desorption, diffusion, wellbore storage, and skin effects. The model delineates seven distinct flow regimes, providing insights into gas migration processes and the factors controlling production. Sensitivity analyses reveal that desorption plays a critical role under low-pressure and low-production conditions, significantly enhancing gas transfer rates from the matrix to the fracture network and contributing to overall production. Monte Carlo simulations further highlight the variability in pressure responses under different input conditions, offering a comprehensive understanding of the model’s behavior in complex reservoir environments. These findings advance the characterization of shale gas flow dynamics and inform the optimization of production strategies. Full article
(This article belongs to the Special Issue Advanced Technology in Unconventional Resource Development)
Show Figures

Figure 1

Figure 1
<p>Physical model of trilinear flow for fractured horizontal well in shale gas reservoir.</p>
Full article ">Figure 2
<p>Flowchart of pressure transient analysis for fractured shale gas wells using trilinear flow model. (<b>a</b>) Model pre−processing: Dimensionless variables are defined, physical continuity across the three flow regions is analyzed, and a mathematical model is established, followed by Laplace transformation. (<b>b</b>) The model is solved in the Laplace domain, and inverse transformation is applied to retrieve solutions in the real space domain. (<b>c</b>) The forward model is employed to perform Monte Carlo simulations, enabling quantification of input parameter uncertainties and comprehensive sensitivity analysis.</p>
Full article ">Figure 3
<p>(<b>a</b>) A comparison of the pressure response characteristic curves of a trilinear flow model for a shale gas, dual − medium, multi − stage fracturing horizontal well and a trilinear flow model for non-desorption diffusion flow; (<b>b</b>) Flow regime division: Stage I: Wellbore storage stage; Stage II: Transition Stage Post-Wellbore Storage; Stage III: Bilinear Flow Between Fractures and Reservoir; Stage IV: Reservoir Linear Flow Stage; Stage V: Desorption and Diffusion Processes in Inner Region Matrix System; Stage VI: Desorption and Diffusion Processes in Outer Region Matrix System; Stage VII: Pseudo-steady-state flow stage of the entire system.</p>
Full article ">Figure 4
<p>Uncertainty analysis of pressure transients using Monte Carlo simulations.</p>
Full article ">Figure 5
<p>Effect of desorption coefficients on pressure response characteristic curves: (<b>a</b>) desorption coefficients in outer zone; (<b>b</b>) desorption coefficients in inner zone.</p>
Full article ">Figure 6
<p>Effect of elastic storage ratio on pressure response characteristic curves: (<b>a</b>) elastic storage ratio in outer zone; (<b>b</b>) elastic storage ratio in inner zone.</p>
Full article ">Figure 7
<p>Effect of channeling coefficient on pressure response characteristic curves: (<b>a</b>) channeling coefficient in outer zone; (<b>b</b>) channeling coefficient in inner zone.</p>
Full article ">
19 pages, 2247 KiB  
Article
Diode Laser Absorption Spectroscopy and DSMC Calculations for the Determination of Species-Specific Diffusion Coefficients of a CO2-N2O Gas Mixture in the Transition Gas Regime
by Kannan Munusamy, Harald Kleine and Sean O’Byrne
Spectrosc. J. 2024, 2(4), 287-305; https://doi.org/10.3390/spectroscj2040017 - 25 Nov 2024
Viewed by 383
Abstract
Multicomponent gas mixture diffusion in a microscale confined flow in the transition gas regime at Knudsen numbers (Kn) above 0.1 has potential engineering applications in gas-phase microfluidics. Although the calculation of the diffusion coefficient accounts for the influence of the concentration of other [...] Read more.
Multicomponent gas mixture diffusion in a microscale confined flow in the transition gas regime at Knudsen numbers (Kn) above 0.1 has potential engineering applications in gas-phase microfluidics. Although the calculation of the diffusion coefficient accounts for the influence of the concentration of other species in a multicomponent gas mixture, the higher rate of gas-wall collision at 0.1 < Kn ≤ 10 introduces additional complications not predicted by conventional calculation methods. Thus, simultaneous measurement of diffusion coefficients for multiple gas species ensures accurate estimation of the diffusion coefficient of a particular species that includes the effect of interactions with other species and wall surface conditions in a multicomponent gas mixture at Kn > 0.1. However, most experimental methods for measuring the diffusion coefficient are not species-specific and therefore cannot directly differentiate between the species diffusing in a gas mixture. Thus, this paper demonstrates a new experiment methodology consisting of a two-bulb diffusion configuration accompanied by a tunable diode laser absorption spectroscopy detection technique for species-specific, in-situ, simultaneous measurement of the effective diffusion coefficient for a CO2-N2O gas mixture in the transition gas regime. The experimental results are compared against direct simulation Monte Carlo calculations and the Bosanquet approximation showing a deviation that has not been reported in the literature before. Full article
(This article belongs to the Special Issue Feature Papers in Spectroscopy Journal)
Show Figures

Figure 1

Figure 1
<p>Schematic arrangement of the two-bulb (TB) diffusion configuration with a hollow-core photonic crystal fiber (HCPCF). The schematic shows a typical direct absorption measurement using the tunable diode laser absorption spectroscopy (TDLAS) technique, shown for a single sawtooth modulation of the laser diode current [<a href="#B41-spectroscj-02-00017" class="html-bibr">41</a>]. Figure is not drawn to scale, and the size of gas particles is exaggerated.</p>
Full article ">Figure 2
<p>(<b>a</b>) Non-absorption regions surrounding two absorption lines selected for the baseline fit interpolation. The rotational transitions are P39e for N<sub>2</sub>O and R31e and R30f for CO<sub>2</sub>. (<b>b</b>) CO<sub>2</sub> and N<sub>2</sub>O absorbance spectrum. (<b>c</b>) Wavenumber interpolation using absorption line peaks. (<b>d</b>) Change in absorbance spectrum during diffusion process shown at different times t = 10 s, 100 s and 900 s.</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>) Non-absorption regions surrounding two absorption lines selected for the baseline fit interpolation. The rotational transitions are P39e for N<sub>2</sub>O and R31e and R30f for CO<sub>2</sub>. (<b>b</b>) CO<sub>2</sub> and N<sub>2</sub>O absorbance spectrum. (<b>c</b>) Wavenumber interpolation using absorption line peaks. (<b>d</b>) Change in absorbance spectrum during diffusion process shown at different times t = 10 s, 100 s and 900 s.</p>
Full article ">Figure 3
<p>Experimental integrated absorbance ∫A<sub>CO<sub>2</sub></sub> and ∫A<sub>N<sub>2</sub>O</sub> and corresponding experimental data fit using the Levenberg-Marquardt non-linear least-squares curve fitting algorithm.</p>
Full article ">Figure 4
<p>Normalised integrated absorbance fit ∫A<sub>CO<sub>2</sub></sub> and ∫A<sub>N<sub>2</sub>O</sub> of CO<sub>2</sub>-N<sub>2</sub>O gas mixture diffusion at Kn = 0.2, 0.3 and 0.6 in the transition gas regime.</p>
Full article ">Figure 5
<p>(<b>a</b>) Cylindrical surface inside the Cartesian co-ordinates simulation box representing experimental diffusion process inside the fiber, and (<b>b</b>) A sample mesh in the <span class="html-italic">yz</span> plane at <span class="html-italic">x</span> = 0 and in the <span class="html-italic">xz</span> plane at <span class="html-italic">y</span> = 0.</p>
Full article ">Figure 6
<p>(<b>a</b>) Initiation and (<b>b</b>) Steady state of DSMC simulation of CO<sub>2</sub>-N<sub>2</sub>O gas mixture diffusion for the simulation length <span class="html-italic">L</span><sub>1</sub> at Kn = 0.6.</p>
Full article ">Figure 7
<p>Change in DSMC computed mole fractions of <math display="inline"><semantics> <msub> <mi>χ</mi> <mrow> <mi>C</mi> <msub> <mi>O</mi> <mn>2</mn> </msub> </mrow> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>χ</mi> <mrow> <msub> <mi>N</mi> <mn>2</mn> </msub> <mi>O</mi> </mrow> </msub> </semantics></math> with time of CO<sub>2</sub>-N<sub>2</sub>O gas mixture diffusion at Kn = 0.6 for TMAC = 1.0.</p>
Full article ">Figure 8
<p>Levenberg-Marquardt nonlinear least-squares error fit for the change in mole fraction <math display="inline"><semantics> <msub> <mi>χ</mi> <mrow> <mi>C</mi> <msub> <mi>O</mi> <mn>2</mn> </msub> </mrow> </msub> </semantics></math> with time of CO<sub>2</sub>-N<sub>2</sub>O gas mixture diffusion at Kn = 0.6 for TMAC = 1.0.</p>
Full article ">Figure 9
<p>The measured, DSMC calculated and Bosanquet approximation predicted <span class="html-italic">D<sub>eff</sub></span> values of CO<sub>2</sub>-N<sub>2</sub>O gas mixture at a few Knudsen numbers in the transition gas regime.</p>
Full article ">
Back to TopTop