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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (89)

Search Parameters:
Keywords = Rayleigh beams

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 1679 KiB  
Article
Vibration Analysis of a Tetra-Layered FGM Cylindrical Shell Using Ring Support
by Asra Ayub, Naveed Hussain, Ahmad N. Al-Kenani and Madiha Ghamkhar
Mathematics 2025, 13(1), 155; https://doi.org/10.3390/math13010155 - 3 Jan 2025
Viewed by 289
Abstract
In the present study, the vibration characteristics of a cylindrical shell (CS) made up of four layers are investigated. The ring is placed in the axial direction of a four-layered functionally graded material (FGM) cylindrical shell. The layers are made of functionally graded [...] Read more.
In the present study, the vibration characteristics of a cylindrical shell (CS) made up of four layers are investigated. The ring is placed in the axial direction of a four-layered functionally graded material (FGM) cylindrical shell. The layers are made of functionally graded material (FGM). The materials used are stainless steel, aluminum, zirconia, and nickel. The frequency equations are derived by employing Sander’s shell theory and the Rayleigh–Ritz (RR) mathematical technique. Vibration characteristics of functionally graded materials have been investigated using polynomial volume fraction law for all FGM layers. The characteristic beam functions have been used to determine the axial model dependency. The natural frequencies are obtained with simply supported boundary conditions by using MATLAB software. Several analogical assessments of shell frequencies have also been conducted to confirm the accuracy and dependability of the current technique. Full article
Show Figures

Figure 1

Figure 1
<p>Flow Chart showing research methodology.</p>
Full article ">Figure 2
<p>CS structure of a four-layered FGM with ring support.</p>
Full article ">Figure 3
<p>Natural frequency (NF) variation with varying power exponent laws in opposition to circumferential wave number for four-layered CS without ring support, when (<span class="html-italic">m</span> = 3, <span class="html-italic">R</span> = 1, <span class="html-italic">N</span> = 3, <span class="html-italic">d</span> = 0.1, <span class="html-italic">L/R</span> = 24, and <span class="html-italic">H</span> = 0.05 <span class="html-italic">m</span>, 0.07 <span class="html-italic">m</span>).</p>
Full article ">Figure 4
<p>Natural frequency (NF) variation with varying power exponent laws in opposition to circumferential wave number for the four-layered cylindrical shell, when (<span class="html-italic">m</span> = 2, <span class="html-italic">d</span> = 0.5, <span class="html-italic">L/R</span> = 20 <span class="html-italic">m</span>, and <span class="html-italic">H</span> = 0.02 <span class="html-italic">m</span>).</p>
Full article ">Figure 5
<p>Natural frequency (NF) variation with varying power exponent laws in opposition to the circumferential wave number for four-layered CS (<span class="html-italic">m</span> = 1, <span class="html-italic">H</span> = 0.006 <span class="html-italic">m</span>, <span class="html-italic">N</span> = 2, <span class="html-italic">R</span> = 1, <span class="html-italic">L/R</span> = 10 <span class="html-italic">m</span> to 30 <span class="html-italic">m</span> and <span class="html-italic">d</span> = 0.2).</p>
Full article ">Figure 6
<p>Natural frequency (NF) variation with ring support ‘<span class="html-italic">d</span>’ at different <span class="html-italic">L/R</span> ratios for four-layered CS without ring support, when (<span class="html-italic">m</span> = 1, <span class="html-italic">R</span> = 1, <span class="html-italic">N</span> = 1, <span class="html-italic">L/R</span> = 30 <span class="html-italic">m</span>, and <span class="html-italic">H</span> = 0.04 <span class="html-italic">m</span>).</p>
Full article ">
13 pages, 396 KiB  
Article
Direct Acceleration of an Electron Beam with a Radially Polarized Long-Wave Infrared Laser
by William H. Li, Igor V. Pogorelsky and Mark A. Palmer
Photonics 2024, 11(11), 1066; https://doi.org/10.3390/photonics11111066 - 14 Nov 2024
Viewed by 646
Abstract
Direct laser acceleration with radially polarized lasers is an intriguing variant of laser-based particle acceleration that has the potential of offering GeV/cm-level energy while avoiding the instabilities and complex beam dynamics associated with plasma wakefield accelerators. A major limiting factor is the difficulty [...] Read more.
Direct laser acceleration with radially polarized lasers is an intriguing variant of laser-based particle acceleration that has the potential of offering GeV/cm-level energy while avoiding the instabilities and complex beam dynamics associated with plasma wakefield accelerators. A major limiting factor is the difficulty of generating high-power radially polarized beams. In this paper, we propose the use of CO2-based long-wave infrared (LWIR) lasers as a driver for direct laser acceleration, as the polarization insensitivity of the gain medium allows a radially polarized beam to be amplified. Additionally, the larger waist sizes, Rayleigh lengths, and pulse lengths associated with the long wavelength could improve the injection efficiency of the electron beam. By comparing acceleration simulations using a near-infrared laser and an LWIR laser, we show that the injection efficiency is indeed improved by up to an order of magnitude with the longer wavelength. Furthermore, we show that even sub-TW peak powers with an LWIR laser can provide MeV-level energy gains. Thus, radially polarized LWIR lasers show significant promise as a driver of a direct laser-driven demonstration accelerator. Full article
(This article belongs to the Special Issue High Power Lasers: Technology and Applications)
Show Figures

Figure 1

Figure 1
<p>Comparison of Ti:Sa and CO<sub>2</sub> direct laser acceleration efficiency. Column titles refer to the electron beam diameter. The term “high-energy fraction" refers to the fraction of the beam that has gained at least 10% of the theoretical gain limit and is used here as a measure of acceleration efficiency. The high-energy fraction of the CO<sub>2</sub> laser is at least 2 times higher than for the Ti:Sa laser at all simulated powers and electron beam diameters, and is almost always 5–10 times higher. The average energy gain is correspondingly higher for the CO<sub>2</sub> laser as well.</p>
Full article ">Figure 2
<p>Energy distributions for some selected peak powers and electron beam diameters. Plots in the same row have the same electron beam diameter and plots in the same column have the same laser peak power. It is clear that, for the Ti:Sa laser, a large portion of the electron beam essentially does not participate in the acceleration interaction, while for the CO<sub>2</sub> laser, the majority of the beam is accelerated to some degree.</p>
Full article ">Figure 3
<p>Energy distribution of a 2 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m initial diameter electron beam after interacting with a 10 PW Ti:Sa with <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>23</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m and <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>86</mn> </mrow> </semantics></math> fs. The electrons experience essentially no acceleration (note that the units on the x-axis are electron volts, not megaelectron volts).</p>
Full article ">Figure 4
<p>Final beam parameters of an accelerated electron beam with an initial energy of 7 MeV. (<b>a</b>) Electron energy distribution. (<b>b</b>) Electron beam divergence. (<b>c</b>–<b>f</b>) Electron beam parameters as a function of beam fraction (defined in the text). The peak energy gain is around a MeV for both laser powers used. The 0.5 TW laser, while providing around half the energy gain, allows for better beam quality.</p>
Full article ">Figure 5
<p>Final beam parameters of an accelerated electron beam with an initial energy of 20 MeV. (<b>a</b>) Electron energy distribution. (<b>b</b>) Electron beam divergence. (<b>c</b>–<b>f</b>) Electron beam parameters as a function of beam fraction (defined in the text). Compared to <a href="#photonics-11-01066-f004" class="html-fig">Figure 4</a>, we see substantially improved acceleration and beam quality, and the 1 TW laser provides around 3 times the peak energy gain as the 0.5 TW laser.</p>
Full article ">Figure A1
<p>Comparison of the simulations conducted in [<a href="#B15-photonics-11-01066" class="html-bibr">15</a>] and a simulated Ti:Sa and CO<sub>2</sub> laser in ASTRA. The laser parameters are listed in the text. We see good agreement between all three sets of simulations, and, notably, the energy gain of the Ti:Sa laser and the CO<sub>2</sub> laser are identical, as predicted.</p>
Full article ">Figure A2
<p>Comparison of a linearly polarized beam and a radially polarized beam. The linearly polarized beam shows noticeably lower energy gain and much higher divergence. The plot of energy evolution corresponds to the particle with the highest final energy. Inset: zoomed-in view of energy oscillations near the focus.</p>
Full article ">
11 pages, 4528 KiB  
Article
Random Raman Lasing in Diode-Pumped Multi-Mode Graded-Index Fiber with Femtosecond Laser-Inscribed Random Refractive Index Structures of Various Shapes
by Alexey G. Kuznetsov, Zhibzema E. Munkueva, Alexandr V. Dostovalov, Alexey Y. Kokhanovskiy, Polina A. Elizarova, Ilya N. Nemov, Alexandr A. Revyakin, Denis S. Kharenko and Sergey A. Babin
Photonics 2024, 11(10), 981; https://doi.org/10.3390/photonics11100981 - 18 Oct 2024
Viewed by 679
Abstract
Diode-pumped multi-mode graded-index (GRIN) fiber Raman lasers provide prominent brightness enhancement both in linear and half-open cavities with random distributed feedback via natural Rayleigh backscattering. Femtosecond laser-inscribed random refractive index structures allow for the sufficient reduction in the Raman threshold by means of [...] Read more.
Diode-pumped multi-mode graded-index (GRIN) fiber Raman lasers provide prominent brightness enhancement both in linear and half-open cavities with random distributed feedback via natural Rayleigh backscattering. Femtosecond laser-inscribed random refractive index structures allow for the sufficient reduction in the Raman threshold by means of Rayleigh backscattering signal enhancement by +50 + 66 dB relative to the intrinsic fiber level. At the same time, they offer an opportunity to generate Stokes beams with a shape close to fundamental transverse mode (LP01), as well as to select higher-order modes such as LP11 with a near-1D longitudinal random structure shifted off the fiber axis. Further development of the inscription technology includes the fabrication of 3D ring-shaped random structures using a spatial light modulator (SLM) in a 100/140 μm GRIN multi-mode fiber. This allows for the generation of a multi-mode diode-pumped GRIN fiber random Raman laser at 976 nm with a ring-shaped output beam at a relatively low pumping threshold (~160 W), demonstrated for the first time to our knowledge. Full article
(This article belongs to the Special Issue Advancements in Fiber Lasers and Their Applications)
Show Figures

Figure 1

Figure 1
<p>Artificial random reflectors of different types fs-inscribed inside the GRIN fiber core: (<b>a</b>) 1D in-line point reflector written along the fiber axis by the direct P-b-P technique, (<b>b</b>) similar 1D in-line reflector shifted off the axis; (<b>c</b>) 3D ring reflector written by SLM-assisted L-b-L technique, meaning that circular lines of overlapping points are inscribed in different planes with the average distance between the planes Δ<span class="html-italic">L</span> and integral length of the structure <span class="html-italic">L</span>.</p>
Full article ">Figure 2
<p>Optical scheme for SLM-assisted writing of random reflective structures using a 4f system with a focus length of 20 cm.</p>
Full article ">Figure 3
<p>Rayleigh backscattering level (<b>a</b>) and reflection spectrum (<b>b</b>) of ring-shaped random structure with length <span class="html-italic">L</span> = 2 mm inscribed in GRIN fiber.</p>
Full article ">Figure 4
<p>Scheme of the Raman fiber laser with a random distributed reflector.</p>
Full article ">Figure 5
<p>Output Stokes power together with a residual pump (<b>a</b>), spectra at different input pump powers (<b>b</b>) and output beam quality and intensity profile (color corresponds to intensity) in the waist shown in the inset, (<b>c</b>) of the MM RFL with an OC in-line random reflector (<span class="html-italic">L</span> = 120 mm, Δ<span class="html-italic">L</span> = 25 µm).</p>
Full article ">Figure 6
<p>Output Stokes power together with a residual pump (<b>a</b>), spectra at different input pump powers (<b>b</b>) and output beam quality and intensity profile (color corresponds to intensity) in the waist shown in the inset (<b>c</b>) of the MM RFL with an OC in-line random reflector (<span class="html-italic">L</span> = 60 mm, Δ<span class="html-italic">L</span> = 50 µm).</p>
Full article ">Figure 7
<p>Output Stokes power together with a residual pump (<b>a</b>), spectra at different input pump powers (<b>b</b>) and output beam quality and intensity profile (color corresponds to intensity) in the waist shown in the inset (<b>c</b>) of the MM RFL with an OC in-line random reflector (<span class="html-italic">L</span> = 2 × 120 mm, Δ<span class="html-italic">L</span> = 25 µm) with a relative shift Δ<span class="html-italic">y</span> ~ 3 μm.</p>
Full article ">Figure 8
<p>(<b>a</b>) Output power of laser with different samples of 3D random distributed reflectors (OC). (<b>b</b>) Generated Stokes power shown in larger scale.</p>
Full article ">Figure 9
<p>Output laser spectra in comparison with HR FBG reflectance spectra for different 3D random reflectors.</p>
Full article ">Figure 10
<p>Measured beam quality parameter M<sup>2</sup> for lasers with different random reflectors at maximum RFL power (from left to right): 2.75, 3, 3.2 μJ. Inset: beam intensity (marked by different colors) profile in the waist.</p>
Full article ">Figure 11
<p>Beam intensity profile at 5.3 W output power captured in the plane of the fiber end face (OC 3.2 μJ).</p>
Full article ">
11 pages, 3614 KiB  
Article
Theoretical Study on Transverse Mode Instability in Raman Fiber Amplifiers Considering Mode Excitation
by Shanmin Huang, Xiulu Hao, Haobo Li, Chenchen Fan, Xiao Chen, Tianfu Yao, Liangjin Huang and Pu Zhou
Micromachines 2024, 15(10), 1237; https://doi.org/10.3390/mi15101237 - 7 Oct 2024
Viewed by 999
Abstract
Raman fiber lasers (RFLs), which are based on the stimulated Raman scattering effect, generate laser beams and offer distinct advantages such as flexibility in wavelength, low quantum defects, and absence from photo-darkening. However, as the power of the RFLs increases, heat generation emerges [...] Read more.
Raman fiber lasers (RFLs), which are based on the stimulated Raman scattering effect, generate laser beams and offer distinct advantages such as flexibility in wavelength, low quantum defects, and absence from photo-darkening. However, as the power of the RFLs increases, heat generation emerges as a critical constraint on further power scaling. This escalating thermal load might result in transverse mode instability (TMI), thereby posing a significant challenge to the development of RFLs. In this work, a static model of the TMI effect in a high-power Raman fiber amplifier based on stimulated thermal Rayleigh scattering is established considering higher-order mode excitation. The variations of TMI threshold power with different seed power levels, fundamental mode purities, higher-order mode losses, and fiber lengths are investigated, while a TMI threshold formula with fundamental mode pumping is derived. This work will enrich the theoretical model of TMI and extend its application scope in TMI mitigation strategies, providing guidance for understanding and suppressing TMI in the RFLs. Full article
(This article belongs to the Special Issue High Power Fiber Laser Technology)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the RFA structure.</p>
Full article ">Figure 2
<p>The change in TMI threshold and output signal LP<sub>01</sub> power with varying seed purities. (<b>a</b>) When the LP<sub>01</sub> power in seed is fixed at 5 W; (<b>b</b>) When the seed power is fixed at 50 W.</p>
Full article ">Figure 3
<p>The TMI threshold under different seed power.</p>
Full article ">Figure 4
<p>The TMI threshold under different pump purities.</p>
Full article ">Figure 5
<p>(<b>a</b>) The TMI threshold and (<b>b</b>) the LP<sub>01</sub> content varies with different seed and pump purities.</p>
Full article ">Figure 6
<p>The pump power, total signal power, and LP<sub>01</sub> power of the signal under different LP<sub>11</sub> mode losses when TMI occurs.</p>
Full article ">Figure 7
<p>(<b>a</b>) Power evolution in RFA; (<b>b</b>) Power evolution and proportion of Stokes frequency-shifted LP<sub>11</sub> mode at different positions along the fiber under a LP<sub>11</sub> mode loss of 4.34 dB/m.</p>
Full article ">Figure 8
<p>(<b>a</b>) Total output signal power and LP<sub>01</sub> power of different fiber lengths; (<b>b</b>) LP<sub>01</sub> mode purity at TMI thresholds of different fiber lengths.</p>
Full article ">
87 pages, 15297 KiB  
Review
Analytic Theory of Seven Classes of Fractional Vibrations Based on Elementary Functions: A Tutorial Review
by Ming Li
Symmetry 2024, 16(9), 1202; https://doi.org/10.3390/sym16091202 - 12 Sep 2024
Viewed by 597
Abstract
This paper conducts a tutorial review of the analytic theory of seven classes of fractional vibrations based on elementary functions. We discuss the classification of seven classes of fractional vibrations and introduce the problem statements. Then, the analytic theory of class VI fractional [...] Read more.
This paper conducts a tutorial review of the analytic theory of seven classes of fractional vibrations based on elementary functions. We discuss the classification of seven classes of fractional vibrations and introduce the problem statements. Then, the analytic theory of class VI fractional vibrators is given. The analytic theories of fractional vibrators from class I to class V and class VII are, respectively, represented. Furthermore, seven analytic expressions of frequency bandwidth of seven classes of fractional vibrators are newly introduced in this paper. Four analytic expressions of sinusoidal responses to fractional vibrators from class IV to VII by using elementary functions are also newly reported in this paper. The analytical expressions of responses (free, impulse, step, and sinusoidal) are first reported in this research. We dissert three applications of the analytic theory of fractional vibrations: (1) analytical expression of the forced response to a damped multi-fractional Euler–Bernoulli beam; (2) analytical expressions of power spectrum density (PSD) and cross-PSD responses to seven classes of fractional vibrators under the excitation with the Pierson and Moskowitz spectrum, which are newly introduced in this paper; and (3) a mathematical explanation of the Rayleigh damping assumption. Full article
Show Figures

Figure 1

Figure 1
<p>Plots of <italic>x</italic><sub>6</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>β</italic>, <italic>λ</italic>) = (1.9, 1.9, 0.3) (solid line), (2.1, 1.8, 0.4) (dotted line), and (2.2, 1.7, 0.5) (dashed line).</p>
Full article ">Figure 2
<p>Plots of <italic>h</italic><sub>6</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>β</italic>, <italic>λ</italic>) = (1.9, 1.9, 0.3) (solid line), (2.1, 1.8, 0.4) (dotted line), and (2.2, 1.7, 0.5) (dashed line).</p>
Full article ">Figure 3
<p>Plots of <italic>g</italic><sub>6</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>β</italic>, <italic>λ</italic>) = (1.9, 1.9, 0.3) (solid line), (2.1, 1.8, 0.4) (dotted line), and (2.2, 1.7, 0.5) (dashed line).</p>
Full article ">Figure 4
<p>Plots of <italic>x</italic><sub>6zs</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>β</italic>, <italic>λ</italic>) = (1.9, 1.9, 0.3) (solid line), (2.1, 1.8, 0.4) (dotted line), and (2.2, 1.7, 0.5) (dashed line).</p>
Full article ">Figure 5
<p>Plots of |<italic>H</italic><sub>6</sub>(<italic>ω</italic>)| and <italic>φ</italic><sub>6</sub>(<italic>ω</italic>) for <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>β</italic>, <italic>λ</italic>) = (1.9, 1.0, 0) (solid line), (2.1, 0.8, 0.1) (dotted line), and (2.5, 0.5, 0.6) (dashed line). (<bold>a</bold>) |<italic>H</italic><sub>6</sub>(<italic>ω</italic>)|; (<bold>b</bold>) <italic>φ</italic><sub>6</sub>(<italic>ω</italic>); (<bold>c</bold>) log<italic>φ</italic><sub>6</sub>(<italic>ω</italic>).</p>
Full article ">Figure 6
<p>Plots of <italic>x</italic><sub>1</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when <italic>α</italic> = 1.5 (solid line), 1.7 (dotted line), and 1.9 (dashed line).</p>
Full article ">Figure 7
<p>Plots of <italic>h</italic><sub>1</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when <italic>α</italic> = 1.5 (solid line), 1.7 (dotted line), and 1.9 (dashed line).</p>
Full article ">Figure 8
<p>Plots of <italic>g</italic><sub>1</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when <italic>α</italic> = 1.5 (solid line), 1.7 (dotted line), and 1.9 (dashed line).</p>
Full article ">Figure 9
<p>Plots of <italic>x</italic><sub>1zs</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when <italic>α</italic> = 1.5 (solid line), 1.7 (dotted line), and 1.9 (dashed line).</p>
Full article ">Figure 10
<p>Plots of |<italic>H</italic><sub>1</sub>(<italic>ω</italic>)| and <italic>φ</italic><sub>1</sub>(<italic>ω</italic>) for <italic>m</italic> = 1 and <italic>k</italic> = 1 when <italic>α</italic> = 1.7 (solid line), 1.9 (dotted line), and 2.2 (dashed line). (<bold>a</bold>) |<italic>H</italic><sub>1</sub>(<italic>ω</italic>)|; (<bold>b</bold>) log|<italic>H</italic><sub>1</sub>(<italic>ω</italic>)|; (<bold>c</bold>) <italic>φ</italic><sub>1</sub>(<italic>ω</italic>).</p>
Full article ">Figure 11
<p>Plots of <italic>x</italic><sub>2</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when <italic>β</italic> = 0.9 (solid line), 1.4 (dotted line), and 1.9 (dashed line).</p>
Full article ">Figure 12
<p>Plots of <italic>h</italic><sub>2</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when <italic>β</italic> = 0.9 (solid line), 1.4 (dotted line), and 1.9 (dashed line).</p>
Full article ">Figure 13
<p>Plots of <italic>g</italic><sub>2</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when <italic>β</italic> = 0.9 (solid line), 1.4 (dotted line), and 1.9 (dashed line).</p>
Full article ">Figure 14
<p>Plots of <italic>x</italic><sub>2zs</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when <italic>β</italic> = 0.9 (solid line), 1.4 (dotted line), and 1.9 (dashed line).</p>
Full article ">Figure 15
<p>Plots of |<italic>H</italic><sub>2</sub>(<italic>ω</italic>)| and <italic>φ</italic><sub>2</sub>(<italic>ω</italic>) for <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when <italic>β</italic> = 0.9 (solid line), 1.5 (dotted line), and 1.9 (dashed line). (<bold>a</bold>) |<italic>H</italic><sub>2</sub>(<italic>ω</italic>)|; (<bold>b</bold>) log|<italic>H</italic><sub>2</sub>(<italic>ω</italic>)|; (<bold>c</bold>). <italic>φ</italic><sub>2</sub>(<italic>ω</italic>).</p>
Full article ">Figure 16
<p>Plots of <italic>x</italic><sub>3</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>β</italic>) = (1.3, 1.9) (solid line), (1.6, 1.8) (dotted line), and (1.9, 1.7) (dashed line).</p>
Full article ">Figure 17
<p>Plots of <italic>h</italic><sub>3</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>β</italic>) = (1.3, 1.9) (solid line), (1.6, 1.8) (dotted line), and (1.9, 1.7) (dashed line).</p>
Full article ">Figure 18
<p>Plots of <italic>g</italic><sub>3</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>β</italic>) = (1.3, 1.9) (solid line), (1.6, 1.8) (dotted line), and (1.9, 1.7) (dashed line).</p>
Full article ">Figure 19
<p>Plots of <italic>x</italic><sub>3zs</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>β</italic>) = (1.3, 1.9) (solid line), (1.6, 1.8) (dotted line), and (1.9, 1.7) (dashed line).</p>
Full article ">Figure 20
<p>Plots of |<italic>H</italic><sub>3</sub>(<italic>ω</italic>)| and <italic>φ</italic><sub>3</sub>(<italic>ω</italic>) for <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>β</italic>) = (1.3, 1.9) (solid line), (1.6, 1.8) (dotted line), and (1.9, 1.7) (dashed line). (<bold>a</bold>) |<italic>H</italic><sub>3</sub>(<italic>ω</italic>)|; (<bold>b</bold>) log|<italic>H</italic><sub>3</sub>(<italic>ω</italic>)|; (<bold>c</bold>) <italic>φ</italic><sub>3</sub>(<italic>ω</italic>); (<bold>d</bold>) log<italic>φ</italic><sub>3</sub>(<italic>ω</italic>).</p>
Full article ">Figure 21
<p>Plots of <italic>x</italic><sub>4</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>λ</italic>) = (1.3, 0.2) (solid line), (1.6, 0.4) (dotted line), and (1.9, 0.6) (dashed line).</p>
Full article ">Figure 22
<p>Plots of <italic>h</italic><sub>4</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>λ</italic>) = (1.3, 0.2) (solid line), (1.6, 0.4) (dotted line), and (1.9, 0.6) (dashed line).</p>
Full article ">Figure 23
<p>Plots of <italic>g</italic><sub>4</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>λ</italic>) = (1.3, 0.2) (solid line), (1.6, 0.4) (dotted line), and (1.9, 0.6) (dashed line).</p>
Full article ">Figure 24
<p>Plots of <italic>x</italic><sub>4zs</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>λ</italic>) = (1.3, 0.2) (solid line), (1.6, 0.5) (dotted line), and (1.9, 0.8) (dashed line).</p>
Full article ">Figure 25
<p>Plots of |<italic>H</italic><sub>4</sub>(<italic>ω</italic>)| and <italic>φ</italic><sub>4</sub>(<italic>ω</italic>) for <italic>m</italic> = 1 and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>λ</italic>) = (1.9, 0.1) (solid line), (2.1, 0.2) (dotted line), and (2.5, 0.3) (dashed line). (<bold>a</bold>) |<italic>H</italic><sub>4</sub>(<italic>ω</italic>)|; (<bold>b</bold>) log|<italic>H</italic><sub>4</sub>(<italic>ω</italic>)|; (<bold>c</bold>) <italic>φ</italic><sub>4</sub>(<italic>ω</italic>); (<bold>d</bold>) log<italic>φ</italic><sub>4</sub>(<italic>ω</italic>).</p>
Full article ">Figure 25 Cont.
<p>Plots of |<italic>H</italic><sub>4</sub>(<italic>ω</italic>)| and <italic>φ</italic><sub>4</sub>(<italic>ω</italic>) for <italic>m</italic> = 1 and <italic>k</italic> = 1 when (<italic>α</italic>, <italic>λ</italic>) = (1.9, 0.1) (solid line), (2.1, 0.2) (dotted line), and (2.5, 0.3) (dashed line). (<bold>a</bold>) |<italic>H</italic><sub>4</sub>(<italic>ω</italic>)|; (<bold>b</bold>) log|<italic>H</italic><sub>4</sub>(<italic>ω</italic>)|; (<bold>c</bold>) <italic>φ</italic><sub>4</sub>(<italic>ω</italic>); (<bold>d</bold>) log<italic>φ</italic><sub>4</sub>(<italic>ω</italic>).</p>
Full article ">Figure 26
<p>Plots of <italic>x</italic><sub>5</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when <italic>λ</italic> = 0.3 (solid line), 0.4 (dot line), and 0.5 (dash line).</p>
Full article ">Figure 27
<p>Plots of <italic>h</italic><sub>5</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when <italic>λ</italic> = 0.3 (solid line), 0.4 (dotted line), and 0.5 (dashed line).</p>
Full article ">Figure 28
<p>Plots of <italic>g</italic><sub>5</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when <italic>λ</italic> = 0.3 (solid line), 0.4 (dotted line), and 0.5 (dashed line).</p>
Full article ">Figure 29
<p>Plots of <italic>x</italic><sub>5zs</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, and <italic>k</italic> = 1 when <italic>λ</italic> = 0.3 (solid line), 0.4 (dotted line), and 0.5 (dashed line).</p>
Full article ">Figure 30
<p>Plots of |<italic>H</italic><sub>5</sub>(<italic>ω</italic>)| and <italic>φ</italic><sub>5</sub>(<italic>ω</italic>) for <italic>m</italic> = 1 and <italic>k</italic> = 1 when <italic>λ</italic> = 0.1 (solid line), 0.2 (dotted line), and 0.3 (dashed line). (<bold>a</bold>) |<italic>H</italic><sub>5</sub>(<italic>ω</italic>)|; (<bold>b</bold>) log|<italic>H</italic><sub>5</sub>(<italic>ω</italic>)|; (<bold>c</bold>) <italic>φ</italic><sub>5</sub>(<italic>ω</italic>).</p>
Full article ">Figure 30 Cont.
<p>Plots of |<italic>H</italic><sub>5</sub>(<italic>ω</italic>)| and <italic>φ</italic><sub>5</sub>(<italic>ω</italic>) for <italic>m</italic> = 1 and <italic>k</italic> = 1 when <italic>λ</italic> = 0.1 (solid line), 0.2 (dotted line), and 0.3 (dashed line). (<bold>a</bold>) |<italic>H</italic><sub>5</sub>(<italic>ω</italic>)|; (<bold>b</bold>) log|<italic>H</italic><sub>5</sub>(<italic>ω</italic>)|; (<bold>c</bold>) <italic>φ</italic><sub>5</sub>(<italic>ω</italic>).</p>
Full article ">Figure 31
<p>Plots of <italic>x</italic><sub>5</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>β</italic>, <italic>λ</italic>) = (1.9, 0.3) (solid line), (1.8, 0.4) (dotted line), and (1.7, 0.5) (dashed line).</p>
Full article ">Figure 32
<p>Plots of <italic>h</italic><sub>7</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>β</italic>, <italic>λ</italic>) = (1.9, 0.3) (solid line), (1.8, 0.4) (dotted line), and (1.7, 0.5) (dashed line).</p>
Full article ">Figure 33
<p>Plots of <italic>g</italic><sub>7</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>β</italic>, <italic>λ</italic>) = (1.9, 0.3) (solid line), (1.8, 0.4) (dotted line), and (1.7, 0.5) (dashed line).</p>
Full article ">Figure 34
<p>Plots of <italic>x</italic><sub>7zs</sub>(<italic>t</italic>) for <italic>ω</italic> = 1.3, <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>β</italic>, <italic>λ</italic>) = (1.9, 0.3) (solid line), (1.8, 0.4) (dotted line), and (1.7, 0.5) (dashed line).</p>
Full article ">Figure 35
<p>Plots of |<italic>H</italic><sub>7</sub>(<italic>ω</italic>)| and <italic>φ</italic><sub>7</sub>(<italic>ω</italic>) for <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>β</italic>, <italic>λ</italic>) = (1.9, 0.05) (solid line), (1.8, 0.10) (dotted line), and (1.7, 0.15) (dashed line). (<bold>a</bold>) |<italic>H</italic><sub>7</sub>(<italic>ω</italic>)|; (<bold>b</bold>) log|<italic>H</italic><sub>7</sub>(<italic>ω</italic>)|; (<bold>c</bold>) <italic>φ</italic><sub>7</sub>(<italic>ω</italic>); (<bold>d</bold>) log<italic>φ</italic><sub>7</sub>(<italic>ω</italic>).</p>
Full article ">Figure 35 Cont.
<p>Plots of |<italic>H</italic><sub>7</sub>(<italic>ω</italic>)| and <italic>φ</italic><sub>7</sub>(<italic>ω</italic>) for <italic>m</italic> = 1, <italic>c</italic> = 0.2, and <italic>k</italic> = 1 when (<italic>β</italic>, <italic>λ</italic>) = (1.9, 0.05) (solid line), (1.8, 0.10) (dotted line), and (1.7, 0.15) (dashed line). (<bold>a</bold>) |<italic>H</italic><sub>7</sub>(<italic>ω</italic>)|; (<bold>b</bold>) log|<italic>H</italic><sub>7</sub>(<italic>ω</italic>)|; (<bold>c</bold>) <italic>φ</italic><sub>7</sub>(<italic>ω</italic>); (<bold>d</bold>) log<italic>φ</italic><sub>7</sub>(<italic>ω</italic>).</p>
Full article ">Figure 36
<p>Plots of <italic>w</italic><sub>1</sub>(<italic>x</italic>, <italic>t</italic>) = w1 of a free-free damped Euler–Bernoulli beam when <italic>f<sub>j</sub></italic>(<italic>t</italic>) = <italic>u</italic>(<italic>t</italic>) and <italic>l</italic> = 10 for <italic>ω</italic> = 0.0314, <italic>a</italic> = 1, <italic>b</italic> = 1, <italic>E</italic> = 1, <italic>x</italic> = 0, …, 10, and <italic>t</italic> = 0, …, 60. (<bold>a</bold>) w1 for <italic>α</italic> = 1.8; <italic>β</italic> = <italic>λ</italic> = 1.2. (<bold>b</bold>) w1 for <italic>α</italic> = 1.8, <italic>β</italic> = 1.2, and <italic>λ</italic> = 0.1.</p>
Full article ">Figure 37
<p>Plot of P-M spectrum.</p>
Full article ">Figure 38
<p>Plots of |<italic>H</italic><sub>1</sub>(<italic>ω</italic>)|<sup>2</sup> with <italic>α</italic> = 1.8 (solid), 2.4 (dotted), and 2.8 (dashed) when <italic>m</italic> = 1 and <italic>k</italic> = 36. (<bold>a</bold>) |<italic>H</italic><sub>1</sub>(<italic>ω</italic>)|<sup>2</sup>; (<bold>b</bold>) log|<italic>H</italic><sub>1</sub>(<italic>ω</italic>)|<sup>2</sup>.</p>
Full article ">Figure 39
<p>Plots of response PSD <italic>S<sub>xx</sub></italic><sub>1</sub>(<italic>ω</italic>) with <italic>α</italic> = 1.8 (solid), 2.4 (dotted), and 2.8 (dashed) when <italic>m</italic> = 1, <italic>k</italic> = 36, and <italic>V</italic> = 15. (<bold>a</bold>) <italic>S<sub>xx</sub></italic><sub>1</sub>(<italic>ω</italic>); (<bold>b</bold>) log<italic>S<sub>xx</sub></italic><sub>1</sub>(<italic>ω</italic>).</p>
Full article ">Figure 40
<p>Resonance when with <italic>α</italic> = 2, <italic>m</italic> = 1, <italic>k</italic> = 36, and <italic>V</italic> = 15. (<bold>a</bold>) Plot of log|<italic>H</italic><sub>1</sub>(<italic>ω</italic>)|<sup>2</sup>; (<bold>b</bold>) plot of response PSD log<italic>S<sub>xx</sub></italic><sub>1</sub>(<italic>ω</italic>).</p>
Full article ">Figure 41
<p>Illustrations of cross-PSD response <italic>S<sub>fx</sub></italic><sub>1</sub>(<italic>ω</italic>) with <italic>α</italic> = 1.8 (solid), 2.4 (dotted), and 2.8 (dashed) when <italic>m</italic> = 1, <italic>k</italic> = 36, and <italic>V</italic> = 15. (<bold>a</bold>) |<italic>S<sub>fx</sub></italic><sub>1</sub>(<italic>ω</italic>)|; (<bold>b</bold>) log|<italic>S<sub>fx</sub></italic><sub>1</sub>(<italic>ω</italic>)|; (<bold>c</bold>) Phase of <italic>S<sub>fx</sub></italic><sub>1</sub>(<italic>ω</italic>). (<bold>d</bold>). Phase of <italic>S<sub>fx</sub></italic><sub>1</sub>(<italic>ω</italic>) in log scale.</p>
Full article ">Figure 42
<p>Plots of random series of driven signal and response signals when <italic>m</italic> = 1, <italic>k</italic> = 36, and <italic>V</italic> = 15: (<bold>a</bold>) Driven signal. (<bold>b</bold>) Response <italic>x</italic><sub>1</sub>(<italic>t</italic>) for <italic>α</italic> = 1.8. (<bold>c</bold>) Response <italic>x</italic><sub>1</sub>(<italic>t</italic>) for <italic>α</italic> = 2.4. (<bold>d</bold>) Response <italic>x</italic><sub>1</sub>(<italic>t</italic>) for <italic>α</italic> = 2.4 in log scale. (<bold>e</bold>) Response <italic>x</italic><sub>1</sub>(<italic>t</italic>) for <italic>α</italic> = 2.8. (<bold>f</bold>) Response <italic>x</italic><sub>1</sub>(<italic>t</italic>) for <italic>α</italic> = 2.8 in log scale.</p>
Full article ">Figure 42 Cont.
<p>Plots of random series of driven signal and response signals when <italic>m</italic> = 1, <italic>k</italic> = 36, and <italic>V</italic> = 15: (<bold>a</bold>) Driven signal. (<bold>b</bold>) Response <italic>x</italic><sub>1</sub>(<italic>t</italic>) for <italic>α</italic> = 1.8. (<bold>c</bold>) Response <italic>x</italic><sub>1</sub>(<italic>t</italic>) for <italic>α</italic> = 2.4. (<bold>d</bold>) Response <italic>x</italic><sub>1</sub>(<italic>t</italic>) for <italic>α</italic> = 2.4 in log scale. (<bold>e</bold>) Response <italic>x</italic><sub>1</sub>(<italic>t</italic>) for <italic>α</italic> = 2.8. (<bold>f</bold>) Response <italic>x</italic><sub>1</sub>(<italic>t</italic>) for <italic>α</italic> = 2.8 in log scale.</p>
Full article ">Figure 43
<p>Plots of |<italic>H</italic><sub>2</sub>(<italic>ω</italic>)|<sup>2</sup> (log) with <italic>β</italic> = 0.4 (solid) and 1.8 (dotted) when <italic>m</italic> = 1, <italic>c</italic> = 0.1, and <italic>k</italic> = 36.</p>
Full article ">Figure 44
<p>Plots of PSD response <italic>S<sub>xx</sub></italic><sub>2</sub>(<italic>ω</italic>) with <italic>β</italic> = 1.8 when <italic>m</italic> = 1, <italic>c</italic> = 0.1, <italic>k</italic> = 36, and <italic>V</italic> = 15: (<bold>a</bold>) <italic>S<sub>xx</sub></italic><sub>2</sub>(<italic>ω</italic>). (<bold>b</bold>) log<italic>S<sub>xx</sub></italic><sub>2</sub>(<italic>ω</italic>).</p>
Full article ">Figure 45
<p>Resonance when with <italic>β</italic> = 1, <italic>m</italic> = 1, <italic>c</italic> = 0.1, <italic>k</italic> = 36, and <italic>V</italic> = 15: (<bold>a</bold>) log|<italic>H</italic><sub>2</sub>(<italic>ω</italic>)|<sup>2</sup>. (<bold>b</bold>) log<italic>S<sub>xx</sub></italic><sub>2</sub>(<italic>ω</italic>).</p>
Full article ">Figure 46
<p>Plots of cross-PSD response <italic>S<sub>fx</sub></italic><sub>2</sub>(<italic>ω</italic>) with <italic>β</italic> = 0.4 (solid) and 1.8 (dot) when <italic>m</italic> = 1, <italic>c</italic> = 0.1, <italic>k</italic> = 36, and <italic>V</italic> = 15: (<bold>a</bold>) log|<italic>S<sub>fx</sub></italic><sub>2</sub>(<italic>ω</italic>)|. (<bold>b</bold>) Phase of <italic>S<sub>fx</sub></italic><sub>2</sub>(<italic>ω</italic>) in log scale.</p>
Full article ">Figure 47
<p>Simulated response series <italic>x</italic><sub>2</sub>(<italic>t</italic>) when <italic>m</italic> = 1, <italic>c</italic> = 0.1, <italic>k</italic> = 36, and <italic>V</italic> = 15: (<bold>a</bold>) <italic>x</italic><sub>2</sub>(<italic>t</italic>) for <italic>β</italic> = 0.4. (<bold>b</bold>) <italic>x</italic><sub>2</sub>(<italic>t</italic>) for <italic>β</italic> = 1. (<bold>c</bold>) <italic>x</italic><sub>2</sub>(<italic>t</italic>) for <italic>β</italic> = 1.4. (<bold>d</bold>) <italic>x</italic><sub>2</sub>(<italic>t</italic>) for <italic>β</italic> = 1.8.</p>
Full article ">Figure 48
<p>Plots of |<italic>H</italic><sub>3</sub>(<italic>ω</italic>)|<sup>2</sup> when <italic>m</italic> = 1, <italic>c</italic> = 0.1, and <italic>k</italic> = 36 for (<italic>α</italic>, <italic>β</italic>) = (1.6, 0.8) (solid), (1.6, 1.8) (dotted), (2.5, 0.8) (dashed), (2.8, 1.8) (dashed–dotted). (<bold>a</bold>) |<italic>H</italic><sub>3</sub>(<italic>ω</italic>)|<sup>2</sup>; (<bold>b</bold>) log|<italic>H</italic><sub>3</sub>(<italic>ω</italic>)|<sup>2</sup>.</p>
Full article ">Figure 49
<p>Plots of response PSD <italic>S<sub>xx</sub></italic><sub>3</sub>(<italic>ω</italic>) when <italic>m</italic> = 1, <italic>c</italic> = 0.1, <italic>k</italic> = 36, and <italic>V</italic> = 15 for (<italic>α</italic>, <italic>β</italic>) = (1.6, 0.8) (solid), (1.6, 1.8) (dotted), (2.5, 0.8) (dashed), and (2.8, 1.8) (dashed–dotted). (<bold>a</bold>) <italic>S<sub>xx</sub></italic><sub>3</sub>(<italic>ω</italic>); (<bold>b</bold>) log<italic>S<sub>xx</sub></italic><sub>3</sub>(<italic>ω</italic>).</p>
Full article ">Figure 50
<p>Plots of cross-PSD response <italic>S<sub>fx</sub></italic><sub>3</sub>(<italic>ω</italic>) when <italic>m</italic> = 1, <italic>c</italic> = 0.1, <italic>k</italic> = 36, and <italic>V</italic> = 15 for (<italic>α</italic>, <italic>β</italic>) = (1.6, 0.8) (solid), (1.6, 1.8) (dotted), (2.5, 0.8) (dashed), and (2.8, 1.8) (dashed–dotted). (<bold>a</bold>) log|<italic>S<sub>fx</sub></italic><sub>3</sub>(<italic>ω</italic>)|; (<bold>b</bold>) Phase of |<italic>S<sub>fx</sub></italic><sub>3</sub>(<italic>ω</italic>)|.</p>
Full article ">Figure 51
<p>Plots of |<italic>H</italic><sub>4</sub>(<italic>ω</italic>)|<sup>2</sup> when <italic>m</italic> = 1, <italic>c</italic> = 0, and <italic>k</italic> = 36 for (<italic>α</italic>, <italic>λ</italic>) = (1.6, 0.2) (solid), (1.6, 0.4) (dotted), (2.5, 0.2) (dashed), and (2.5, 0.4) (dashed–dotted). (<bold>a</bold>) |<italic>H</italic><sub>4</sub>(<italic>ω</italic>)|<sup>2</sup>; (<bold>b</bold>) log|<italic>H</italic><sub>4</sub>(<italic>ω</italic>)|<sup>2</sup>.</p>
Full article ">Figure 52
<p>Plots of response PSD <italic>S<sub>xx</sub></italic><sub>4</sub>(<italic>ω</italic>) when <italic>m</italic> = 1, <italic>c</italic> = 0, <italic>k</italic> = 36, and <italic>V</italic> = 15 for (<italic>α</italic>, <italic>λ</italic>) = (1.6, 0.2) (solid), (1.6, 0.4) (dotted), (2.5, 0.2) (dashed), and (2.5, 0.4) (dashed–dotted). (<bold>a</bold>) <italic>S<sub>xx</sub></italic><sub>4</sub>(<italic>ω</italic>); (<bold>b</bold>) log<italic>S<sub>xx</sub></italic><sub>4</sub>(<italic>ω</italic>).</p>
Full article ">Figure 53
<p>Plots of cross-PSD response <italic>S<sub>fx</sub></italic><sub>4</sub>(<italic>ω</italic>) when <italic>m</italic> = 1, <italic>c</italic> = 0, <italic>k</italic> = 36, and <italic>V</italic> = 15 for (<italic>α</italic>, <italic>λ</italic>) = (1.6, 0.2) (solid), (1.6, 0.4) (dotted), (2.5, 0.2) (dashed), and (2.5, 0.4) (dashed–dotted). (<bold>a</bold>) log|<italic>S<sub>fx</sub></italic><sub>4</sub>(<italic>ω</italic>)|; (<bold>b</bold>) phase of |<italic>S<sub>fx</sub></italic><sub>4</sub>(<italic>ω</italic>)|.</p>
Full article ">Figure 54
<p>Plots of |<italic>H</italic><sub>5</sub>(<italic>ω</italic>)|<sup>2</sup> (log) when <italic>m</italic> = 1, <italic>c</italic> = 0, and <italic>k</italic> = 36 for <italic>λ</italic> = 0.2 (solid), 0.4 (dotted), 0.6 (dashed), and 0.8 (dashed–dotted).</p>
Full article ">Figure 55
<p>Plots of response PSD <italic>S<sub>xx</sub></italic><sub>5</sub>(<italic>ω</italic>) when <italic>m</italic> = 1, <italic>c</italic> = 0, <italic>k</italic> = 36, and <italic>V</italic> = 15 for <italic>λ</italic> = 0.2 (solid), 0.4 (dotted), 0.6 (dashed), and 0.8 (dashed–dotted): (<bold>a</bold>) <italic>S<sub>xx</sub></italic><sub>5</sub>(<italic>ω</italic>); (<bold>b</bold>) log<italic>S<sub>xx</sub></italic><sub>5</sub>(<italic>ω</italic>).</p>
Full article ">Figure 56
<p>Plots of cross-PSD response <italic>S<sub>fx</sub></italic><sub>5</sub>(<italic>ω</italic>) when <italic>m</italic> = 1, <italic>c</italic> = 0, <italic>k</italic> = 36, and <italic>V</italic> = 15 for <italic>λ</italic> = 0.2 (solid), 0.4 (dotted), 0.6) (dashed), and 0.8 (dashed–dotted): (<bold>a</bold>) log|<italic>S<sub>fx</sub></italic><sub>5</sub>(<italic>ω</italic>)|; (<bold>b</bold>) phase of |<italic>S<sub>fx</sub></italic><sub>5</sub>(<italic>ω</italic>)|.</p>
Full article ">Figure 57
<p>Plots of |<italic>H</italic><sub>6</sub>(<italic>ω</italic>)|<sup>2</sup> (log) when <italic>m</italic> = 1, <italic>c</italic> = 0.1, and <italic>k</italic> = 36 for (<italic>α</italic>, <italic>β</italic>, <italic>λ</italic>) = (1.6, 0.8, 0.2) (solid), (1.6, 1.8, 0.4) (dotted), (2.5, 0.4, 0.2) (dashed), and (2.5, 0.8, 0.4) (dashed–dotted).</p>
Full article ">Figure 58
<p>Plots of response PSD <italic>S<sub>xx</sub></italic><sub>6</sub>(<italic>ω</italic>) in log scale when <italic>m</italic> = 1, <italic>c</italic> = 0.1, <italic>k</italic> = 36, and <italic>V</italic> = 15 for (<italic>α</italic>, <italic>β</italic>, <italic>λ</italic>) = (1.6, 0.8, 0.2) (solid), (1.6, 1.8, 0.4) (dotted), (2.5, 0.4, 0.2) (dashed), and (2.5, 0.8, 0.4) (dashed–dotted).</p>
Full article ">Figure 59
<p>Plots of cross-PSD response <italic>S<sub>fx</sub></italic><sub>6</sub>(<italic>ω</italic>) when <italic>m</italic> = 1, <italic>c</italic> = 0.1, <italic>k</italic> = 36, and <italic>V</italic> = 15 for (<italic>α</italic>, <italic>β</italic>, <italic>λ</italic>) = (1.6, 0.8, 0.2) (solid), (1.6, 1.8, 0.4) (dotted), (2.5, 0.4, 0.2) (dashed), and (2.5, 0.8, 0.4) (dashed–dotted): (<bold>a</bold>) log|<italic>S<sub>fx</sub></italic><sub>6</sub>(<italic>ω</italic>)|; (<bold>b</bold>) phase of |<italic>S<sub>fx</sub></italic><sub>6</sub>(<italic>ω</italic>)|.</p>
Full article ">Figure 60
<p>Plots of |<italic>H</italic><sub>7</sub>(<italic>ω</italic>)|<sup>2</sup> (log) when <italic>m</italic> = 1, <italic>c</italic> = 0.1, and <italic>k</italic> = 36, for (<italic>β</italic>, <italic>λ</italic>) = (0.5, 0.2) (solid), (0.5, 0.3) (dot), (1.5, 0.4) (dash), (1.5, 0.6) (dash dot).</p>
Full article ">Figure 61
<p>Plots of response PSD <italic>S<sub>xx</sub></italic><sub>7</sub>(<italic>ω</italic>) in log scale when <italic>m</italic> = 1, <italic>c</italic> = 0.1, <italic>k</italic> = 36, and <italic>V</italic> = 15 for (<italic>β</italic>, <italic>λ</italic>) = (0.5, 0.2) (solid), (0.5, 0.3) (dot), (1.5, 0.4) (dash), (1.5, 0.6) (dash dot).</p>
Full article ">Figure 62
<p>Plots of cross-PSD response <italic>S<sub>fx</sub></italic><sub>7</sub>(<italic>ω</italic>) when <italic>m</italic> = 1, <italic>c</italic> = 0.1, <italic>k</italic> = 36, and <italic>V</italic> = 15 for (<italic>β</italic>, <italic>λ</italic>) = (0.5, 0.2) (solid), (0.5, 0.3) (dot), (1.5, 0.4) (dash), (1.5, 0.6) (dash dot). (<bold>a</bold>). log|<italic>S<sub>fx</sub></italic><sub>7</sub>(<italic>ω</italic>)|. (<bold>b</bold>). Phase of |<italic>S<sub>fx</sub></italic><sub>7</sub>(<italic>ω</italic>)| in log scale.</p>
Full article ">
11 pages, 1759 KiB  
Article
Gold Nanoparticles at a Liquid Interface: Towards a Soft Nonlinear Metasurface
by Delphine Schaming, Anthony Maurice, Frédéric Gumy, Micheál D. Scanlon, Christian Jonin, Hubert H. Girault and Pierre-François Brevet
Photonics 2024, 11(9), 789; https://doi.org/10.3390/photonics11090789 - 23 Aug 2024
Cited by 1 | Viewed by 731
Abstract
Optical second-harmonic generation (SHG) is achieved using adsorbed gold nanoparticles (AuNPs) with an average diameter of 16 nm at the aqueous solution–air interface in reflection. A detailed analysis of the depth profile of the SHG intensity detected shows that two contributions appear in [...] Read more.
Optical second-harmonic generation (SHG) is achieved using adsorbed gold nanoparticles (AuNPs) with an average diameter of 16 nm at the aqueous solution–air interface in reflection. A detailed analysis of the depth profile of the SHG intensity detected shows that two contributions appear in the overall signal, one arising from the aqueous solution–air interface that is sensitive to the AuNP surface excess and one arising from the bulk aqueous phase. The latter is an incoherent signal also known as hyper-Rayleigh scattering (HRS). The results agree with those of an analysis involving Gaussian beam propagation optics and a Langmuir-like isotherm. Discrepancies are revealed for the largest AuNP concentrations used and indicate a new route for the design of soft metasurfaces. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental set-up of the aqueous drop at the end of the capillary tube, this itself mounted on a vertical translation stage. The bottom air–aqueous drop interface and volume are illuminated with the fundamental beam of a femtosecond Ti-Sa laser at 810 nm, and the output SHG light is captured with a photomultiplier tube placed behind a monochromator. The light beam is focused and captured through the same objective (see the text for further details). Red line indicates path of the fundamental beam whereas blue line indicates path of harmonic beam.</p>
Full article ">Figure 2
<p>Extinction spectrum for a 2 nM aqueous solution of 16 nm diameter gold nanoparticles (AuNPs), synthesized using Turkevich’s method. Insert: Transmission electron microscopy (TEM) of sample; the scale bar is 20 nm.</p>
Full article ">Figure 3
<p>SHG intensity profile, collected at 405 nm, of the aqueous solution–air interface as the position of the laser focal point was scanned from within and outside of a water droplet in the absence of AuNPs. The disks represent the experimental data and the red line represents the fit to the model (see <a href="#sec3dot3-photonics-11-00789" class="html-sec">Section 3.3</a>). The parameter obtained from the model is <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>; see below.</p>
Full article ">Figure 4
<p>SHG intensity profile, collected at 405 nm, of the AuNP-modified aqueous solution–air interface as the position of the laser focal point was scanned from within and outside of a water droplet containing a 1:4 <span class="html-italic">v</span>/<span class="html-italic">v</span> dilution of a 2 nM aqueous solution measuring 16 nm in diameter AuNPs. At z &gt; 0.40 mm, the laser beam waist was positioned in the bulk aqueous phase, and at z &lt; 0.40 mm, it was located in air. The disks represent the experimental data, and the red line represents the fit to the model (see <a href="#sec3dot3-photonics-11-00789" class="html-sec">Section 3.3</a>). The parameter obtained from the model is <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1.65</mn> </mrow> </semantics></math>; see below.</p>
Full article ">Figure 5
<p>SHG intensity, obtained at 405 nm, of the AuNP-modified aqueous solution–air interface as the position of the laser focal point was scanned in and out of a water droplet containing a 2 nM aqueous solution of 16 nm diameter AuNPs. At <span class="html-italic">z</span> &gt; 0.40 mm, the laser beam waist was positioned in the bulk aqueous phase and at <span class="html-italic">z</span> &lt; 0.40 mm, it was located in air. The disks represent the experimental data, and the red line represents the theoretical model fitting (see <a href="#sec3dot3-photonics-11-00789" class="html-sec">Section 3.3</a>). The parameter obtained from the model is <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>7.9</mn> </mrow> </semantics></math>; see below.</p>
Full article ">Figure 6
<p>Plot of the <span class="html-italic">α</span> parameter as a function of the concentration of the aqueous 16 nm diameter AuNP solution. The disks represent the experimental data, and the red line represents the linear adjustment.</p>
Full article ">
22 pages, 5370 KiB  
Article
A Novel Semi-Active Control Approach for Flexible Structures: Vibration Control through Boundary Conditioning Using Magnetorheological Elastomers
by Jomar Morales and Ramin Sedaghati
Vibration 2024, 7(2), 605-626; https://doi.org/10.3390/vibration7020032 - 18 Jun 2024
Viewed by 762
Abstract
This research study explores an alternative method of vibration control of flexible beam type structures via boundary conditioning using magnetorheological elastomer at the support location. The Rayleigh–Ritz method has been used to formulate dynamic equations of motions of the beam with MRE support [...] Read more.
This research study explores an alternative method of vibration control of flexible beam type structures via boundary conditioning using magnetorheological elastomer at the support location. The Rayleigh–Ritz method has been used to formulate dynamic equations of motions of the beam with MRE support and to extract its natural frequencies and mode shapes. The MRE-based adaptive continuous beam is then converted into an equivalent single-degree-of-freedom system for the purpose of control implementation, assuming that the system’s response is dominated by its fundamental mode. Two different types of control strategies are formulated including proportional–integral–derivative control and on–off control. The performance of controllers is evaluated for three different loading conditions including shock, harmonic, and random vibration excitations. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of Experimental Data and Curve Fitting Representation of Storage and Loss Moduli vs. Magnetic Flux Densities at 2 Hz and 15% Shear Strain [<a href="#B16-vibration-07-00032" class="html-bibr">16</a>].</p>
Full article ">Figure 2
<p>Demagnetization Curve for N52 Permanent Magnet.</p>
Full article ">Figure 3
<p>Electromagnet Layout with Parametric Dimensions.</p>
Full article ">Figure 4
<p>(<b>a</b>) FEMM Results for Input Current of −3 A; (<b>b</b>) FEMM Results for Input Current of 3 A.</p>
Full article ">Figure 5
<p>Comparison of the FEMM Results with the Curve Fitted 4th-Order Polynomial of the Magnetic Flux Density vs. Input Current Curve.</p>
Full article ">Figure 6
<p>(<b>a</b>) Overhang Beam with Spring at Overhang Support; (<b>b</b>) Equivalent Beam Model with MRE in Direct shear.</p>
Full article ">Figure 7
<p>Shear and Bending Moment Shown in Red of (<b>a</b>) Cantilever Beam and (<b>b</b>) Overhanging Beam.</p>
Full article ">Figure 8
<p>Equivalent Single-Degree-of-Freedom System.</p>
Full article ">Figure 9
<p>Curve Fit for Equivalent Stiffness Storage and Loss Components, Equivalent Mass, and Equivalent Damping.</p>
Full article ">Figure 10
<p>Comparison of Transient Response and Input Current between Passive and Semi-active Systems due to different Control Strategies.</p>
Full article ">Figure 11
<p>Steady-State Time Response and Control Current for Different Controllers under Harmonic Input at (<b>a</b>) 5.14 Hz and (<b>b</b>) at 7.06 Hz.</p>
Full article ">Figure 12
<p>Time Response and Controller Input for Random Input over (<b>a</b>) 50 Seconds and (<b>b</b>) 5 Seconds.</p>
Full article ">Figure A1
<p>Static and Dynamic Shear Strain Representations on Deflected MRE.</p>
Full article ">Figure A2
<p>Free-Body Diagram of Proposed Beam Model under Static Loading.</p>
Full article ">
12 pages, 9274 KiB  
Article
Optical Force Effects of Rayleigh Particles by Cylindrical Vector Beams
by Yuting Zhao, Liqiang Zhou, Xiaotong Jiang, Linwei Zhu and Qiang Shi
Nanomaterials 2024, 14(8), 691; https://doi.org/10.3390/nano14080691 - 17 Apr 2024
Cited by 2 | Viewed by 1013
Abstract
High-order cylindrical vector beams possess flexible spatial polarization and exhibit new effects and phenomena that can expand the functionality and enhance the capability of optical systems. However, building a general analytical model for highly focused beams with different polarization orders remains a challenge. [...] Read more.
High-order cylindrical vector beams possess flexible spatial polarization and exhibit new effects and phenomena that can expand the functionality and enhance the capability of optical systems. However, building a general analytical model for highly focused beams with different polarization orders remains a challenge. Here, we elaborately develop the vector theory of high-order cylindrical vector beams in a high numerical aperture focusing system and achieve the vectorial diffraction integrals for describing the tight focusing field with the space-variant distribution of polarization orders within the framework of Richards–Wolf diffraction theory. The analytical formulae include the exact three Cartesian components of electric and magnetic distributions in the tightly focused region. Additionally, utilizing the analytical formulae, we can achieve the gradient force, scattering force, and curl-spin force exerted on Rayleigh particles trapped by high-order cylindrical vector beams. These results are crucial for improving the design and engineering of the tightly focused field by modulating the polarization orders of high-order cylindrical vector beams, particularly for applications such as optical tweezers and optical manipulation. This theoretical analysis also extends to the calculation of complicated optical vortex vector fields and the design of diffractive optical elements with high diffraction efficiency and resolution. Full article
(This article belongs to the Special Issue Advances in Optical Nanomanipulation)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of focusing system geometry.</p>
Full article ">Figure 2
<p>Polarization distributions of incident vector beam with different orders: (<b>a</b>) <span class="html-italic">m</span> = 0, <span class="html-italic">ϕ</span><sub>0</sub> = 0, (<b>b</b>) <span class="html-italic">m</span> = 1, <span class="html-italic">ϕ</span><sub>0</sub> = 0, (<b>c</b>) <span class="html-italic">m</span> = −1, <span class="html-italic">ϕ</span><sub>0</sub> = 0, and (<b>d</b>) <span class="html-italic">m</span> = 2, <span class="html-italic">ϕ</span><sub>0</sub> = 0. Arrows indicate the direction of polarization at the arrow location.</p>
Full article ">Figure 3
<p>The intensity distribution on the focal plane of vector beams <span class="html-italic">E<sub>x</sub></span>, <span class="html-italic">E<sub>y</sub></span>, <span class="html-italic">E<sub>z</sub></span> and <span class="html-italic">E<sub>total</sub></span> with different orders: (<b>a</b>–<b>d</b>) <span class="html-italic">m</span> = 0, (<b>e</b>–<b>h</b>) <span class="html-italic">m</span> = 1, (<b>i</b>–<b>l</b>) <span class="html-italic">m</span> = −1, (<b>m</b>–<b>p</b>) <span class="html-italic">m</span> = 2.</p>
Full article ">Figure 4
<p>The intensity distribution on the focal plane of vector beams <span class="html-italic">H<sub>x</sub></span>, <span class="html-italic">H<sub>y</sub></span>, <span class="html-italic">H<sub>z</sub></span> and <span class="html-italic">H<sub>total</sub></span> with different orders: (<b>a</b>–<b>d</b>) <span class="html-italic">m</span> = 0, (<b>e</b>–<b>h</b>) <span class="html-italic">m</span> = 1, (<b>i</b>–<b>l</b>) <span class="html-italic">m</span> = −1, (<b>m</b>–<b>p</b>) <span class="html-italic">m</span> = 2.</p>
Full article ">Figure 5
<p>The iso-surface intensity distribution of the electric and magnetic fields in the tightly focused region with <span class="html-italic">E</span> = 0.5<span class="html-italic">E<sub>max</sub></span> and <span class="html-italic">H</span> = 0.5<span class="html-italic">H<sub>max</sub></span>.</p>
Full article ">Figure 6
<p>The intensity contour of the electric and magnetic fields in the focal plane for the vector beams: (<b>a</b>,<b>b</b>) m = 0, <span class="html-italic">ϕ</span><sub>0</sub> = π/4, (<b>c</b>,<b>d</b>) m = 0, <span class="html-italic">ϕ</span><sub>0</sub> = π/2, (<b>e</b>,<b>f</b>) m = 1, <span class="html-italic">ϕ</span><sub>0</sub> = π/4, (<b>g</b>,<b>h</b>) m = 1, <span class="html-italic">ϕ</span><sub>0</sub> = π/2, (<b>i</b>,<b>j</b>) m = −1, <span class="html-italic">ϕ</span><sub>0</sub> = π/4, (<b>k</b>,<b>l</b>) m = −1, <span class="html-italic">ϕ</span><sub>0</sub> =π/2, (<b>m</b>,<b>n</b>) m = 2, <span class="html-italic">ϕ</span><sub>0</sub> = π/4, (<b>o</b>,<b>p</b>) m = 2, <span class="html-italic">ϕ</span><sub>0</sub> = π/2. The short lines indicate the polarization distributions.</p>
Full article ">Figure 7
<p>The gradient force distributions produced by highly focused CV beams with <span class="html-italic">m</span> equal to 0, 1, −1, and 2. (<b>a</b>–<b>d</b>) represent the transverse gradient force distributions. (<b>e</b>–<b>h</b>) represent the longitudinal gradient force distributions. Arrows denote the direction and magnitude of the gradient force.</p>
Full article ">Figure 8
<p>The longitudinal scattering force in the focal plane for different orders <span class="html-italic">m</span> of CV beam: (<b>a</b>) <span class="html-italic">m</span> = 0, (<b>b</b>) <span class="html-italic">m</span> = 1, (<b>c</b>) <span class="html-italic">m</span> = −1, (<b>d</b>) <span class="html-italic">m</span> = 2. Arrows denote the direction and magnitude of the scattering force.</p>
Full article ">Figure 9
<p>The longitudinal curl-spin force for different orders <span class="html-italic">m</span> of CV beam: (<b>a</b>) <span class="html-italic">m</span> = 0, (<b>b</b>) <span class="html-italic">m</span> = 1, (<b>c</b>) <span class="html-italic">m</span> = −1, (<b>d</b>) <span class="html-italic">m</span> = 2. Arrows denote the direction and magnitude of the curl-spin force.</p>
Full article ">
38 pages, 1755 KiB  
Article
The Fresnel Approximation and Diffraction of Focused Waves
by Colin J. R. Sheppard
Photonics 2024, 11(4), 346; https://doi.org/10.3390/photonics11040346 - 9 Apr 2024
Viewed by 2328
Abstract
In this paper, diffraction of scalar waves by a screen with a circular aperture is explored, considering the incidence of either a collimated beam or a focused wave, a historical review of the development of the theory is presented, and the introduction of [...] Read more.
In this paper, diffraction of scalar waves by a screen with a circular aperture is explored, considering the incidence of either a collimated beam or a focused wave, a historical review of the development of the theory is presented, and the introduction of the Fresnel approximation is described. For diffraction by a focused wave, the general case is considered for both high numerical aperture and for finite values of the Fresnel number. One aim is to develop a theory based on the use of dimensionless optical coordinates that can help to determined the general behaviour and trends of different system parameters. An important phenomenon, the focal shift effect, is discussed as well. Explicit expressions are provided for focal shift and the peak intensity for different numerical apertures and Fresnel numbers. This is one application where the Rayleigh–Sommerfeld diffraction integrals provide inaccurate results. Full article
(This article belongs to the Special Issue Laser Beam Propagation and Control)
Show Figures

Figure 1

Figure 1
<p>Contours of <span class="html-italic">u</span> (blue) and <span class="html-italic">v</span> (red) for FrA1 for diffraction of a plane wave by a circular aperture.</p>
Full article ">Figure 2
<p>Contours of <span class="html-italic">u</span> (blue) and <span class="html-italic">v</span> (red) for FrA2 for diffraction of a plane wave by a circular aperture.</p>
Full article ">Figure 3
<p>Contours of <span class="html-italic">u</span> (blue) and <span class="html-italic">v</span> (red) for the generalized Fresnel approximation for diffraction of a plane wave by a circular aperture.</p>
Full article ">Figure 4
<p>Contours of <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>=</mo> <mi>k</mi> <mi>z</mi> <mo>+</mo> <msup> <mi>v</mi> <mn>2</mn> </msup> <mo>/</mo> <mn>2</mn> <mi>u</mi> </mrow> </semantics></math> for the generalized Fresnel approximation for diffraction of a plane wave by a circular aperture.</p>
Full article ">Figure 5
<p>Schematic diagram of a spherical wave diffracted by a circular aperture of radius <span class="html-italic">a</span> in an opaque screen at distance <span class="html-italic">d</span> from the geometrical focal point <span class="html-italic">F</span>, with sagitta <math display="inline"><semantics> <mrow> <mi>h</mi> <mo>=</mo> <mi>f</mi> <mo>−</mo> <mi>d</mi> </mrow> </semantics></math>; the angle <math display="inline"><semantics> <mi>α</mi> </semantics></math> is the semi-angular aperture, and <math display="inline"><semantics> <mi>θ</mi> </semantics></math> is the angle subtended at the optical axis by a ray from a general point <span class="html-italic">W</span> on the wavefront to <span class="html-italic">F</span>. The points <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>E</mi> <mn>2</mn> </msub> </mrow> </semantics></math> on the aperture edge correspond to the maximum and minimum distances <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>g</mi> <mn>2</mn> </msub> </mrow> </semantics></math> from the observation point <span class="html-italic">P</span>, with cylindrical coordinates <math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>,</mo> <mi>ϕ</mi> <mo>,</mo> <mi>z</mi> </mrow> </semantics></math> relative to <span class="html-italic">F</span> and distant <math display="inline"><semantics> <msub> <mi>r</mi> <mi>P</mi> </msub> </semantics></math> from <span class="html-italic">F</span> and <span class="html-italic">r</span> from <span class="html-italic">W</span>. The triangle <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>F</mi> <mi>P</mi> <msub> <mi>E</mi> <mn>2</mn> </msub> </mrow> </semantics></math> has an exterior angle <math display="inline"><semantics> <mrow> <mo>∡</mo> <msub> <mi>E</mi> <mn>2</mn> </msub> <mi>F</mi> <mi>A</mi> <mo>=</mo> <msub> <mi>α</mi> <mn>2</mn> </msub> </mrow> </semantics></math> and interior angles <math display="inline"><semantics> <mrow> <mo>∡</mo> <mi>F</mi> <mi>P</mi> <msub> <mi>E</mi> <mn>2</mn> </msub> <mo>=</mo> <msub> <mi>β</mi> <mn>2</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>∡</mo> <mi>F</mi> <msub> <mi>E</mi> <mn>2</mn> </msub> <mi>P</mi> <mo>=</mo> <msub> <mi>δ</mi> <mn>2</mn> </msub> </mrow> </semantics></math>. The angle rays between rays from point <span class="html-italic">W</span> to <span class="html-italic">F</span> and <span class="html-italic">P</span> are <math display="inline"><semantics> <mrow> <mo>∡</mo> <mi>F</mi> <mi>W</mi> <mi>P</mi> <mo>=</mo> <mi>ψ</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>The behaviour of (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>/</mo> <mn>2</mn> <mi>π</mi> <msub> <mi>N</mi> <mi>f</mi> </msub> </mrow> </semantics></math> (from Ks) and (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>u</mi> <mn>0</mn> </msub> <mo>/</mo> <mn>2</mn> <mi>π</mi> <msub> <mi>N</mi> <mi>f</mi> </msub> </mrow> </semantics></math> (from aKs), with <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> for different semi-angular apertures. Lines are coloured red for small <math display="inline"><semantics> <mi>α</mi> </semantics></math>, orange for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, green for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, blue for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, and purple for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. For small <math display="inline"><semantics> <mi>α</mi> </semantics></math>, the behaviour of <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>/</mo> <mn>2</mn> <mi>π</mi> <mi>N</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>u</mi> <mn>0</mn> </msub> <mo>/</mo> <mn>2</mn> <mi>π</mi> <msub> <mi>N</mi> <mi>f</mi> </msub> </mrow> </semantics></math> agrees with that of Li and Wolf [<a href="#B50-photonics-11-00346" class="html-bibr">50</a>]. Along the optical axis, <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>/</mo> <mn>2</mn> <mi>π</mi> <msub> <mi>N</mi> <mi>f</mi> </msub> <mo>=</mo> <mi>Z</mi> <mo>/</mo> <mi>f</mi> <mo>=</mo> <mi>D</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>/</mo> <mi>f</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>, with <span class="html-italic">Z</span> as in Equation (<a href="#FD42-photonics-11-00346" class="html-disp-formula">42</a>) and <span class="html-italic">D</span> as in Equation (<a href="#FD46-photonics-11-00346" class="html-disp-formula">46</a>).</p>
Full article ">Figure 7
<p>The fractional focal shift as a function of the Fresnel number <span class="html-italic">N</span> for different NAs: (<b>a</b>) for the Ks model (Equations (<a href="#FD46-photonics-11-00346" class="html-disp-formula">46</a>) and (<a href="#FD55-photonics-11-00346" class="html-disp-formula">55</a>)) and (<b>b</b>) from the analytic aKs expression in Equation (<a href="#FD59-photonics-11-00346" class="html-disp-formula">59</a>). Lines are coloured red for small <math display="inline"><semantics> <mi>α</mi> </semantics></math>, orange for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, green for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, blue for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, and purple for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. For small <math display="inline"><semantics> <mi>α</mi> </semantics></math>, the behaviour agrees with that of Li and Wolf [<a href="#B50-photonics-11-00346" class="html-bibr">50</a>]. The curves for small <math display="inline"><semantics> <mi>α</mi> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> are very close to each other. The small circles indicate where <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>/</mo> <mi>λ</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math>: for small <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>/</mo> <mi>λ</mi> </mrow> </semantics></math>, the Kirchhoff boundary conditions tend to break down.</p>
Full article ">Figure 8
<p>Contours of constant intensity in the focal region of a focused scalar wave diffracted by a circular aperture in the Debye approximation: (<b>a</b>) for a small semi-angular aperture <math display="inline"><semantics> <mi>α</mi> </semantics></math>, (<b>b</b>) for the nonparaxial Debye case for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>c</b>) for the nonparaxial Debye case for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, and (<b>d</b>) for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, computed from the analytic expression in Equation (<a href="#FD65-photonics-11-00346" class="html-disp-formula">65</a>). The intensity at the focal point is normalized to unity. An analytic plot for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> was first shown in [<a href="#B86-photonics-11-00346" class="html-bibr">86</a>].</p>
Full article ">Figure 9
<p>Contours of <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (in red) and <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (in green for <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> <mo>&lt;</mo> <mn>0</mn> </mrow> </semantics></math> and blue for <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>) for the scaled Debye–Wolf theory for diffraction of a spherical wave by a circular aperture: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. The black contour is for <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>/</mo> <mi>f</mi> <mo>=</mo> <mo>[</mo> <mi>W</mi> <mo>−</mo> <mn>2</mn> <mi>Z</mi> <msup> <mo form="prefix">sin</mo> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mi>α</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> <mo>]</mo> <mo>/</mo> <mi>f</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. The shadow edge is also indicated in black.</p>
Full article ">Figure 10
<p>Contours of <math display="inline"><semantics> <mrow> <mi>W</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (in green for <math display="inline"><semantics> <mrow> <mi>W</mi> <mo>/</mo> <mi>f</mi> <mo>&lt;</mo> <mn>0</mn> </mrow> </semantics></math> and blue for <math display="inline"><semantics> <mrow> <mi>W</mi> <mo>/</mo> <mi>f</mi> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>) for the scaled Debye–Wolf theory for diffraction of a spherical wave by a circular aperture: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. The black contour is for <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>/</mo> <mi>f</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>A magnified view of the contours of <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (in red) and <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (in green for <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> <mo>≤</mo> <mn>0</mn> </mrow> </semantics></math> and blue for <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>) for the scaled Debye–Wolf theory for diffraction of a spherical wave by a circular aperture: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. The shadow edge is indicated in black. The contours for <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> provided by the two alternative expressions in Equation (<a href="#FD74-photonics-11-00346" class="html-disp-formula">74</a>) are almost identical at this scale.</p>
Full article ">Figure 12
<p>The variations in (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>W</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> and (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>/</mo> <mi>f</mi> <mo>=</mo> <mo>[</mo> <mi>W</mi> <mo>−</mo> <mn>2</mn> <mi>Z</mi> <msup> <mo form="prefix">sin</mo> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mi>α</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> <mo>]</mo> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> along the shadow edge. Here, <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mi>P</mi> </msub> <mo>=</mo> <msqrt> <mrow> <msup> <mi>z</mi> <mn>2</mn> </msup> <mo>+</mo> <msup> <mi>ρ</mi> <mn>2</mn> </msup> </mrow> </msqrt> </mrow> </semantics></math>. Lines are coloured orange for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, green for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, blue for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, and purple for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 13
<p>Contours of <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>/</mo> <mi>f</mi> <mo>=</mo> <mo>[</mo> <mi>W</mi> <mo>−</mo> <mn>2</mn> <mi>Z</mi> <msup> <mo form="prefix">sin</mo> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mi>α</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> <mo>]</mo> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> for the scaled Debye–Wolf theory for diffraction of a spherical wave by a circular aperture: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 14
<p>Plots of <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (<b>top row</b>), <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (<b>middle row</b>), and <math display="inline"><semantics> <mrow> <mi>W</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (<b>bottom row</b>) along lines of constant <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> ((<b>left column</b>) <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mo>−</mo> <mn>0.2</mn> <mi>f</mi> </mrow> </semantics></math>, (<b>middle column</b>) <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, and (<b>right column</b>) <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0.2</mn> <mi>f</mi> </mrow> </semantics></math>), shown for different values of <math display="inline"><semantics> <mi>α</mi> </semantics></math> (indicated by colour, orange <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, green <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, blue <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, and purple <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>) Break points are indicated by small circles.</p>
Full article ">Figure 15
<p>Plots showing contours of constant <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (<b>left column: a and c</b>) and <math display="inline"><semantics> <mrow> <mi>W</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (<b>right column: b and d</b>) for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> (<b>top row: a and b</b>) and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> (<b>bottom row: c and d</b>) for the csDW theory. Colours: illuminated region, blue <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>, green <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>&lt;</mo> <mn>0</mn> </mrow> </semantics></math>; shadow region, purple.</p>
Full article ">Figure 16
<p>Plots showing contours of constant <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (in green) (<b>left column: a and c</b>) and <math display="inline"><semantics> <mrow> <mi>W</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (in blue) (<b>right column: b and d</b>) for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> (<b>top row: a and b</b>) and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> (<b>bottom row: c and d</b>) for the asDW theory.</p>
Full article ">Figure 17
<p>Contours of <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (in red), <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>W</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> (in green for <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> <mo>,</mo> <mi>W</mi> <mo>/</mo> <mi>f</mi> <mo>&lt;</mo> <mn>0</mn> </mrow> </semantics></math> and blue for <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> <mo>,</mo> <mi>W</mi> <mo>/</mo> <mi>f</mi> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>) for the pseudo-paraxial theory for diffraction of a spherical wave by a circular aperture: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>W</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>W</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. The black contours in (<b>c</b>,<b>d</b>) are for <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>/</mo> <mi>f</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 18
<p>(<b>a</b>) The variation in <math display="inline"><semantics> <msup> <mi>C</mi> <mn>2</mn> </msup> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> for RSI (red), RSII (blue), and Kp (green). Curves are shown for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> (solid line), <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> (dashed line), and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>75</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> (chained line). The variation in <math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mo>=</mo> <msup> <mrow> <mo>[</mo> <mn>2</mn> <mi>f</mi> <mo>/</mo> <mrow> <mo>(</mo> <mi>f</mi> <mo>+</mo> <mi>z</mi> <mo>+</mo> <mi>g</mi> <mo>)</mo> </mrow> <mo>]</mo> </mrow> <mn>2</mn> </msup> </mrow> </semantics></math> is shown in purple. (<b>b</b>) The variation in <math display="inline"><semantics> <msup> <mrow> <mo>(</mo> <mn>1</mn> <mo>−</mo> <mi>C</mi> <mo>)</mo> </mrow> <mn>2</mn> </msup> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>f</mi> </mrow> </semantics></math> for RSI (red), RSII (blue), and Kp (green). Curves are shown for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> (solid line), <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>60</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> (dashed line), and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>75</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> (chained line).</p>
Full article ">
11 pages, 5767 KiB  
Technical Note
The Impacts of Deformed Fabry–Perot Interferometer Transmission Spectrum on Wind Lidar Measurements
by Ming Zhao, Jianfeng Chen, Chenbo Xie and Lu Li
Remote Sens. 2024, 16(6), 1076; https://doi.org/10.3390/rs16061076 - 19 Mar 2024
Cited by 1 | Viewed by 1050
Abstract
The Fabry–Perot interferometer (FPI) plays a crucial role as the frequency discriminator in the incoherent Doppler wind lidar. However, in the practical receiver system, reflections occurring between optical elements introduce non-normal incident components in the light beams passing through the FPI. This phenomenon [...] Read more.
The Fabry–Perot interferometer (FPI) plays a crucial role as the frequency discriminator in the incoherent Doppler wind lidar. However, in the practical receiver system, reflections occurring between optical elements introduce non-normal incident components in the light beams passing through the FPI. This phenomenon results in the deformation of the FPI transmission spectral lines. Based on that, a theoretical model has been developed to describe the transmission spectrum of the FPI when subjected to obliquely incident light beams with a divergence angle. By appropriately adjusting the model parameters, the simulated transmission spectrum of the FPI edge channels can coincide with the experimentally measured FPI spectral line. Subsequently, the impact of deformations in the transmission spectrum of the two edge channels on wind measurements is evaluated. The first implication is a systematic shift of 30.7 m/s in line-of-sight (LOS) wind velocities. This shift is based on the assumption that the lidar echo is solely backscattered from atmospheric molecules. The second consequence is the inconsistency in the response sensitivities of Doppler frequency shift between Rayleigh signals and Mie signals. As a result, the lidar system fails to fully achieve its initial design objectives, particularly in effectively suppressing interference from Mie signals. The presence of aerosols can introduce a significant error of several meters per second in the measurement of LOS wind velocity. Full article
(This article belongs to the Section Environmental Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Transmission spectrum lines of edge channels under different incident beam divergence angles.</p>
Full article ">Figure 2
<p>(<b>a</b>) The case of beam with divergence angle obliquely incident on FPI. (<b>b</b>) The geometric position of any light ray in the beam.</p>
Full article ">Figure 3
<p>The geometric relationship within the incident surface.</p>
Full article ">Figure 4
<p>The transmission spectrum lines of the FPI correspond to different incident angles.</p>
Full article ">Figure 5
<p>The transmission spectrum lines of the FPI when multiple beams are incident. (<b>a</b>) Spectral components in the model (<b>b</b>) Measured transmission spectrum line, Airy fitting spectrum line, and calculated spectrum line for multiple beam incidence.</p>
Full article ">Figure 6
<p>Principle of Doppler shift determination using a triple channel FPI.</p>
Full article ">Figure 7
<p>The response curves of Rayleigh and Mie signals that derived from the actual FPI transmission spectral line and the designed FPI transmission spectral line.</p>
Full article ">Figure 8
<p>The input horizontal (<b>a</b>) wind velocity and (<b>b</b>) wind direction profiles.</p>
Full article ">Figure 9
<p>The input (solid lines) and retrieved (dashed lines) LOS wind velocity profiles.</p>
Full article ">Figure 10
<p>The system error in the retrieved horizontal (<b>a</b>) wind velocity and (<b>b</b>) wind direction.</p>
Full article ">Figure 11
<p>The designed and actual response sensitivities for Rayleigh and Mie signals.</p>
Full article ">Figure 12
<p>Calculation of LOS wind velocity errors caused by aerosols with different transmission spectral lines. (<b>a</b>) design FPI spectral lines (<b>b</b>) actual FPI spectral lines.</p>
Full article ">
10 pages, 27451 KiB  
Communication
Spin Hall Effect of Two-Index Paraxial Vector Propagation-Invariant Beams
by Victor V. Kotlyar and Alexey A. Kovalev
Photonics 2023, 10(11), 1288; https://doi.org/10.3390/photonics10111288 - 20 Nov 2023
Cited by 2 | Viewed by 1162
Abstract
We investigate a simple paraxial vector beam, which is a coaxial superposition of two single-ringed Laguerre–Gaussian (LG) beams, linearly polarized along the horizontal axis, with topological charges (TC) n and −n, and of two LG beams, linearly polarized along the vertical [...] Read more.
We investigate a simple paraxial vector beam, which is a coaxial superposition of two single-ringed Laguerre–Gaussian (LG) beams, linearly polarized along the horizontal axis, with topological charges (TC) n and −n, and of two LG beams, linearly polarized along the vertical axis, with the TCs m and −m. In the initial plane, such a vector beam has zero spin angular momentum (SAM). Upon propagation in free space, such a propagation-invariant beam has still zero SAM at several distances from the waist plane (initial plane). However, we show that at all other distances, the SAM becomes nonzero. The intensity distribution in the cross-section of such a beam has 2m (if m > n) lobes, the maxima of which reside on a circle of a certain radius. The SAM distribution has also several lobes, from 2m till 2(m + n), the centers of which reside on a circle with a radius smaller than that of the maximal-intensity circle. The SAM sign alternates differently: one lobe has a positive SAM, while two neighbor lobes on the circle have a negative SAM, or two neighbor pairs of lobes can have a positive and negative SAM. When passing through a plane with zero SAM, positive and negative SAM lobes are swapped. The maximal SAM value is achieved at a distance smaller than or equal to the Rayleigh distance. Full article
(This article belongs to the Special Issue Structured Light Beams: Science and Applications)
Show Figures

Figure 1

Figure 1
<p>Intensity (<b>a</b>) and SAM density (<b>b</b>) distributions of the light field (2) at a distance <span class="html-italic">z</span> = 2<span class="html-italic">z</span><sub>0</sub> for the following computation parameters: wavelength λ = 0.532 μm, waist radius of the Gaussian envelope <span class="html-italic">w</span><sub>0</sub> = 1 mm, beam orders <span class="html-italic">m</span> = 2 and <span class="html-italic">n</span> = 3, and initial phases <span class="html-italic">α</span> = 0, <span class="html-italic">β</span> = −π/2. Scale mark in each figure denotes 1 mm. Black and yellow color (<b>a</b>) denote zero and maximal intensity. Blue and red color (<b>b</b>) denote positive and negative SAM density. Pink and cyan ellipses (<b>b</b>) denote right- and left-handed elliptic polarization. White lines (<b>b</b>) denote lines with zero SAM density. Dashed circles (<b>a</b>) illustrate the circles with maximal SAM density (green circle) and with maximal intensity of the components <span class="html-italic">E<sub>x</sub></span> and <span class="html-italic">E<sub>y</sub></span> (white circles) obtained using Equations (12) and (13).</p>
Full article ">Figure 2
<p>Intensity (<b>a</b>–<b>d</b>) and SAM density (<b>e</b>–<b>h</b>) distributions of the light field (2) in several transverse planes for the following computation parameters: wavelength λ = 0.532 μm, waist radius of the Gaussian envelope <span class="html-italic">w</span><sub>0</sub> = 1 mm, beams orders <span class="html-italic">m</span> = 7 and <span class="html-italic">n</span> = 14, initial phases <span class="html-italic">α</span> = 0, <span class="html-italic">β</span> = −π/2, and propagation distances <span class="html-italic">z</span> = 0 (<b>a</b>,<b>e</b>), <span class="html-italic">z</span> = <span class="html-italic">z</span><sub>0</sub>/2 (<b>b</b>,<b>f</b>), <span class="html-italic">z</span> = <span class="html-italic">z</span><sub>0</sub> (<b>b</b>,<b>f</b>), and <span class="html-italic">z</span> = 2<span class="html-italic">z</span><sub>0</sub> (<b>b</b>,<b>f</b>). The scale mark in each figure denotes 1 mm. Dashed circles (d,h) illustrate the circles with maximal SAM density (green circle) and with the maximal intensity of the components <span class="html-italic">E<sub>x</sub></span> and <span class="html-italic">E<sub>y</sub></span> (white circles) obtained using Equations (12) and (13). Numbers near the color bar under each figure denote minimal and maximal values.</p>
Full article ">Figure 3
<p>Intensity (<b>a</b>–<b>d</b>) and SAM density (<b>e</b>–<b>h</b>) distributions of the light field (2) in several transverse planes for the following computation parameters: wavelength λ = 0.532 μm, waist radius of the Gaussian envelope <span class="html-italic">w</span><sub>0</sub> = 1 mm, beams orders <span class="html-italic">m</span> = 3 and <span class="html-italic">n</span> = 9, initial phases <span class="html-italic">α</span> = <span class="html-italic">β</span> = 0, and propagation distances <span class="html-italic">z</span> = 0 (<b>a</b>,<b>e</b>), <span class="html-italic">z</span> = <span class="html-italic">z</span><sub>0</sub>/2 (<b>b</b>,<b>f</b>), <span class="html-italic">z</span> = <span class="html-italic">z</span><sub>0</sub> (<b>b</b>,<b>f</b>), and <span class="html-italic">z</span> = 2<span class="html-italic">z</span><sub>0</sub> (<b>b</b>,<b>f</b>). Scale mark in each figure denotes 1 mm. Numbers near the color bar under each figure denote minimal and maximal values.</p>
Full article ">Figure 4
<p>Dependence of the maximal SAM density of the light field (2), with the same parameters as in <a href="#photonics-10-01288-f003" class="html-fig">Figure 3</a>, on the propagation distance <span class="html-italic">z</span>/<span class="html-italic">z</span><sub>0</sub>. Green dots correspond to the maximal SAM densities from <a href="#photonics-10-01288-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 5
<p>Intensity (<b>a</b>,<b>c</b>) and SAM density (<b>b</b>,<b>d</b>) distributions of the light field (2) for the following computation parameters: wavelength λ = 0.532 μm, waist radius of the Gaussian envelope <span class="html-italic">w</span><sub>0</sub> = 1 mm, beams orders <span class="html-italic">m</span> = 6 and <span class="html-italic">n</span> = 14 (<b>a</b>,<b>b</b>), <span class="html-italic">m</span> = 7 and <span class="html-italic">n</span> = 14 (<b>c</b>,<b>d</b>), initial phases <span class="html-italic">α</span> = 0, <span class="html-italic">β</span> = −π/2, and propagation distance <span class="html-italic">z</span> = 2<span class="html-italic">z</span><sub>0</sub>. Scale mark in each figure denotes 1 mm. Numbers near the color bar under each figure denote minimal and maximal values.</p>
Full article ">
27 pages, 13939 KiB  
Article
Free Vibration Analysis of Elastically Restrained Tapered Beams with Concentrated Mass and Axial Force
by Jung Woo Lee
Appl. Sci. 2023, 13(19), 10742; https://doi.org/10.3390/app131910742 - 27 Sep 2023
Cited by 3 | Viewed by 1265
Abstract
This study proposes a new numerical method for the free vibration analysis of elastically restrained tapered Rayleigh beams with concentrated mass and axial force. The beam model had elastic support, concentrated mass at both ends, and axial force at the right end. The [...] Read more.
This study proposes a new numerical method for the free vibration analysis of elastically restrained tapered Rayleigh beams with concentrated mass and axial force. The beam model had elastic support, concentrated mass at both ends, and axial force at the right end. The elastic supports were modeled as translational and rotational springs. The shear force and bending moment were determined under the assumption that the sum of the forces at arbitrary positions and the joint between the beam and elastic supports always becomes zero. Therefore, a frequency determinant is established considering the free-free end condition at both ends, but various boundary conditions were constructed by adjusting the values of the elastic springs in the frequency equation. This assumption simplified the deduction procedure, and the method’s efficiency was demonstrated through various comparisons. In particular, the value of compressive loading at which the first natural frequency vanished was investigated by considering the taper ratio based on the relationship between the elastic support and compressive loading. The analyzed results can be adopted as benchmark solutions for other approaches. The frequency determinant employs the transfer matrix method; however, numerical methods can easily be utilized in other approaches. Full article
(This article belongs to the Section Acoustics and Vibrations)
Show Figures

Figure 1

Figure 1
<p>Geometry of elastically restrained tapered beam with concentrated mass and axial loading.</p>
Full article ">Figure 2
<p>Beam types utilized to consider various conditions: (<b>a</b>) Type A, (<b>b</b>) Type B, (<b>c</b>) Type C, and (<b>d</b>) Type D.</p>
Full article ">Figure 3
<p>Variation of nondimensional natural frequency with respect to increase of spring value.</p>
Full article ">Figure 4
<p>Effects of the concentrated mass on the first three natural frequencies of single tapered beams with the cantilevered end condition: (<b>a</b>) results for entire concentrated masses, and (<b>b</b>) results for <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>m</mi> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math> = 0, 0.1, 0.5, and 1.</p>
Full article ">Figure 5
<p>Effects of the concentrated mass on the first three natural frequencies of double tapered beams with the cantilevered end condition: (<b>a</b>) results for entire concentrated masses, and (<b>b</b>) results for <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>m</mi> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math> = 0, 0.1, 0.5, and 1.</p>
Full article ">Figure 6
<p>Comparison of natural frequencies for <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> = 1 and <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> = 2 when having <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>m</mi> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math> = 0.1 at the arbitrary location.</p>
Full article ">Figure 7
<p>Effects of the concentrated mass and taper ratios on the first three natural frequencies of <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> = 1: Cases A1–A4.</p>
Full article ">Figure 8
<p>Effects of the concentrated mass and taper ratios on the first three natural frequencies of <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> = 1 with respect to Cases A5–A8.</p>
Full article ">Figure 9
<p>Effects of the concentrated mass and taper ratios on the first three natural frequencies of <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> = 1 with respect to Cases A9–A12.</p>
Full article ">Figure 10
<p>Effects of the concentrated mass and taper ratios on the first three natural frequencies of <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> = 2 with respect to Cases A1–A4.</p>
Full article ">Figure 11
<p>Effects of the concentrated mass and taper ratios on the first three natural frequencies of <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> = 2 with respect to Cases A5–A8.</p>
Full article ">Figure 12
<p>Effects of the concentrated mass and taper ratios on the first three natural frequencies of <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> = 2: Cases A9–A12.</p>
Full article ">Figure 13
<p>Comparison of <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> = 1 and <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> = 2 for effects of the concentrated mass on the first three natural frequencies when <span class="html-italic">c</span> = 0.5: (<b>a</b>) Cases A3 and A4, (<b>b</b>) Cases A7 and A8, and (<b>c</b>) Cases A11 and A12.</p>
Full article ">Figure 14
<p>Effects of the axial loading on the first three natural frequencies of elastically restrained tapered beam.</p>
Full article ">Figure 15
<p>Relationships between compressive loadings and values of elastic springs at which the first natural frequency vanished.</p>
Full article ">Figure 16
<p>Variation in the first three mode shapes of tapered beams for Type A.</p>
Full article ">Figure 17
<p>Effects of the elastic supports on the first three mode shapes of uniform beams and single and double tapered beams having <math display="inline"><semantics> <mrow> <mi>c</mi> </mrow> </semantics></math> = 0.5.</p>
Full article ">
26 pages, 980 KiB  
Article
Dynamic Response of an Elastic Tube-like Nanostructure Embedded in a Vibrating Medium and under the Action of Moving Nano-Objects
by Xiaoxia Ma, Mojtaba Roshan, Keivan Kiani and Ali Nikkhoo
Symmetry 2023, 15(10), 1827; https://doi.org/10.3390/sym15101827 - 26 Sep 2023
Cited by 7 | Viewed by 1847
Abstract
In recent years, researchers have looked at how tube-like nanostructures respond to moving loads and masses. However, no one has explored the scenario of a nanostructure embedded in a vibrating medium used for moving nano-objects. In this study, the governing equations of the [...] Read more.
In recent years, researchers have looked at how tube-like nanostructures respond to moving loads and masses. However, no one has explored the scenario of a nanostructure embedded in a vibrating medium used for moving nano-objects. In this study, the governing equations of the problem are methodically derived using the nonlocal elasticity of Eringen as well as the Rayleigh and Reddy–Bickford beam theories. Analytical and numerical solutions are developed for capturing the nonlocal dynamic deflection of the nanostructure based on the moving nanoforce approach (excluding the inertia effect) and the moving nanomass approach (including the inertia effect), respectively. The results predicted by the established models are successfully verified with those of other researchers in some special cases. The results reveal that for low velocities of the moving nano-object in the absence of the medium excitation, the midspan deflection of the simply supported nanotube exhibits an almost symmetric time-history curve; however, by increasing the nano-object velocity or the medium excitation amplitude, such symmetry is violated, mainly due to the lateral inertia of the moving nano-object, as displayed by the corresponding three-dimensional plots. The study addresses the effects of the mass and velocity of the moving nano-object, amplitude, and frequency of the medium excitation, and the lateral and rotational stiffness of the nearby medium in contact with the nanostructure on the maximum dynamic deflection. The achieved results underscore the significance of considering both the inertial effect of the moving nano-object and the shear effect of stocky nanotubes embedded in vibrating media. This research can serve as a strong basis for conducting further investigations into the vibrational properties of more intricate tube-shaped nanosystems that are embedded in a vibrating medium, with the aim of delivering nano-objects. Full article
Show Figures

Figure 1

Figure 1
<p>An embedded continuum-based tube-like nanostructure acted upon by the excitations of both elastic medium and MNO.</p>
Full article ">Figure 2
<p>(<b>a</b>) Plots of <math display="inline"><semantics> <msub> <mi>W</mi> <mrow> <mi>N</mi> <mo>,</mo> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </semantics></math>-<math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mi>b</mi> </msub> <mo>/</mo> <mi>d</mi> </mrow> </semantics></math> (<math display="inline"><semantics> <msub> <mi>r</mi> <mi>m</mi> </msub> </semantics></math> = 0.325 nm, <math display="inline"><semantics> <msub> <mi>ρ</mi> <mi>b</mi> </msub> </semantics></math> = 2300 kg/m<math display="inline"><semantics> <msup> <mrow/> <mn>3</mn> </msup> </semantics></math>, <math display="inline"><semantics> <msub> <mi>t</mi> <mi>b</mi> </msub> </semantics></math> = 0.35 nm, <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.1, <math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> = 0; (∘) <math display="inline"><semantics> <mrow> <msub> <mi>e</mi> <mn>0</mn> </msub> <mi>a</mi> </mrow> </semantics></math> = 0, <math display="inline"><semantics> <mfenced open="(" close=")"> <mo>▵</mo> </mfenced> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>e</mi> <mn>0</mn> </msub> <mi>a</mi> </mrow> </semantics></math> = 1 nm, (⋄) <math display="inline"><semantics> <mrow> <msub> <mi>e</mi> <mn>0</mn> </msub> <mi>a</mi> </mrow> </semantics></math> = 2 nm; (—) Simsek [<a href="#B108-symmetry-15-01827" class="html-bibr">108</a>], and (...) present study). (<b>b</b>) Plots of <math display="inline"><semantics> <msub> <mi>W</mi> <mrow> <mi>N</mi> <mo>,</mo> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </semantics></math>-<math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> (<math display="inline"><semantics> <msub> <mi>r</mi> <mi>m</mi> </msub> </semantics></math> = 3 nm, <math display="inline"><semantics> <msub> <mi>ρ</mi> <mi>b</mi> </msub> </semantics></math> = 2500 kg/m<math display="inline"><semantics> <msup> <mrow/> <mn>3</mn> </msup> </semantics></math>, <math display="inline"><semantics> <msub> <mi>M</mi> <mi>N</mi> </msub> </semantics></math> = 0, <math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> = 0.01<math display="inline"><semantics> <msub> <mi>t</mi> <mi>b</mi> </msub> </semantics></math>, <math display="inline"><semantics> <mi>λ</mi> </semantics></math> = 30; (—) Kiani [<a href="#B103-symmetry-15-01827" class="html-bibr">103</a>], (...) present study; (∘) NRABT, and <math display="inline"><semantics> <mfenced open="(" close=")"> <mo>▵</mo> </mfenced> </semantics></math> NREBT) retained.</p>
Full article ">Figure 3
<p>Plots of <math display="inline"><semantics> <msub> <mi>W</mi> <mrow> <mi>N</mi> <mo>,</mo> <mi>m</mi> <mi>i</mi> <mi>d</mi> </mrow> </msub> </semantics></math>-<math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>/</mo> <msub> <mi>τ</mi> <mi>f</mi> </msub> </mrow> </semantics></math> based on the NREBT: (<b>a</b>) <math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> = 0, (<b>b</b>) <math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> = <math display="inline"><semantics> <mrow> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> <msub> <mi>l</mi> <mi>b</mi> </msub> </mrow> </semantics></math>, and (<b>c</b>) <math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>2</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> <msub> <mi>l</mi> <mi>b</mi> </msub> </mrow> </semantics></math> (<math display="inline"><semantics> <mi>β</mi> </semantics></math> = 0.6, <math display="inline"><semantics> <msub> <mi>M</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>r</mi> <mi>R</mi> </msubsup> </semantics></math> = <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>t</mi> <mi>R</mi> </msubsup> </semantics></math> = 10, <math display="inline"><semantics> <mi>λ</mi> </semantics></math> = 20; (...) MNFA, (—) MNMA; (□) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.03, and <math display="inline"><semantics> <mfenced open="(" close=")"> <mo>▵</mo> </mfenced> </semantics></math> <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, and (⋄) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.6).</p>
Full article ">Figure 4
<p>Three-dimensional plots of <math display="inline"><semantics> <msub> <mi>W</mi> <mrow> <mi>N</mi> <mo>,</mo> <mi>m</mi> <mi>i</mi> <mi>d</mi> </mrow> </msub> </semantics></math>-<math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math>-<math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>/</mo> <msub> <mi>τ</mi> <mi>f</mi> </msub> </mrow> </semantics></math> based on the NRABT: (<b>a</b>) in the absence of medium excitation (<math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> = 0); (<b>b</b>) in the presence of medium excitation (<math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> = <math display="inline"><semantics> <mrow> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> <msub> <mi>l</mi> <mi>b</mi> </msub> </mrow> </semantics></math>); (<math display="inline"><semantics> <mi>β</mi> </semantics></math> = 0.6, <math display="inline"><semantics> <msub> <mi>M</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>r</mi> <mi>R</mi> </msubsup> </semantics></math> = <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>t</mi> <mi>R</mi> </msubsup> </semantics></math> = 10, and <math display="inline"><semantics> <mi>λ</mi> </semantics></math> = 20).</p>
Full article ">Figure 5
<p>Plots of <math display="inline"><semantics> <msub> <mi>W</mi> <mrow> <mi>N</mi> <mo>,</mo> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </semantics></math>-<math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> for various MNO’s speeds: (<b>a</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, (<b>b</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.6, and (<b>c</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.9; ((∘) MNFA, <math display="inline"><semantics> <mfenced open="(" close=")"> <mo>▵</mo> </mfenced> </semantics></math> MNMA; (...) NRABT, and (—) NREBT; <math display="inline"><semantics> <mi>β</mi> </semantics></math> = 0.6, <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mi>g</mi> </msub> <mo>/</mo> <msub> <mi>l</mi> <mi>b</mi> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> </semantics></math>, <math display="inline"><semantics> <msub> <mi>M</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>t</mi> <mi>R</mi> </msubsup> </semantics></math> = 10, and <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>r</mi> <mi>R</mi> </msubsup> </semantics></math> = 10).</p>
Full article ">Figure 6
<p>Plots of <math display="inline"><semantics> <msub> <mi>W</mi> <mrow> <mi>N</mi> <mo>,</mo> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </semantics></math>-<math display="inline"><semantics> <mi>β</mi> </semantics></math> based on the MNMA for various MNO’s speeds: (<b>a</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, (<b>b</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.6, and (<b>c</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.9; ((...) NRABT, (—) NREBT; <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mi>g</mi> </msub> <mo>/</mo> <msub> <mi>l</mi> <mi>b</mi> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> </semantics></math>, <math display="inline"><semantics> <msub> <mi>M</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>t</mi> <mi>R</mi> </msubsup> </semantics></math> = 10, <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>r</mi> <mi>R</mi> </msubsup> </semantics></math> = 10; and <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>β</mi> </mstyle> </semantics></math> = <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mover> <mi>ω</mi> <mo>¯</mo> </mover> <msub> <mi>ω</mi> <mn>1</mn> </msub> </mfrac> </mstyle> </semantics></math>).</p>
Full article ">Figure 7
<p>Plots of <math display="inline"><semantics> <msub> <mi>W</mi> <mrow> <mi>N</mi> <mo>,</mo> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </semantics></math>-MNO’s mass for various MNO’s speeds: (<b>a</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, (<b>b</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.6, and (<b>c</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.9; ((∘) MNFA, <math display="inline"><semantics> <mfenced open="(" close=")"> <mo>▵</mo> </mfenced> </semantics></math> MNMA; (...) NRABT, and (—) NREBT; <math display="inline"><semantics> <mi>β</mi> </semantics></math> = 0.6, <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mi>g</mi> </msub> <mo>/</mo> <msub> <mi>l</mi> <mi>b</mi> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> </semantics></math>, and <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>t</mi> <mi>R</mi> </msubsup> </semantics></math> = 10, <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>r</mi> <mi>R</mi> </msubsup> </semantics></math> = 10).</p>
Full article ">Figure 8
<p>Plots of <math display="inline"><semantics> <msub> <mi>W</mi> <mrow> <mi>N</mi> <mo>,</mo> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </semantics></math>-MNO’s speed for three levels of the excitation amplitude of the elastic medium: (<b>a</b>) <math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> = <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> <msub> <mi>l</mi> <mi>b</mi> </msub> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>2</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> <msub> <mi>l</mi> <mi>b</mi> </msub> </mrow> </semantics></math>, and (<b>c</b>) <math display="inline"><semantics> <msub> <mi>a</mi> <mi>g</mi> </msub> </semantics></math> = <math display="inline"><semantics> <mrow> <mn>3</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> <msub> <mi>l</mi> <mi>b</mi> </msub> </mrow> </semantics></math>; ((∘) MNFA, <math display="inline"><semantics> <mfenced open="(" close=")"> <mo>▵</mo> </mfenced> </semantics></math> MNMA; (...) NRABT, and (—) NREBT; <math display="inline"><semantics> <msub> <mi>M</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, and <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>t</mi> <mi>R</mi> </msubsup> </semantics></math> = <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>r</mi> <mi>R</mi> </msubsup> </semantics></math> = 10).</p>
Full article ">Figure 9
<p>Graphs of <math display="inline"><semantics> <msub> <mi>W</mi> <mrow> <mi>N</mi> <mo>,</mo> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </semantics></math>-<math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>r</mi> <mi>R</mi> </msubsup> </semantics></math> for various MNO’s speeds: (<b>a</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, (<b>b</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.6, and (<b>c</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.9; ((∘) MNFA, <math display="inline"><semantics> <mfenced open="(" close=")"> <mo>▵</mo> </mfenced> </semantics></math> MNMA; (...) NRABT, and (—) NREBT; <math display="inline"><semantics> <mi>β</mi> </semantics></math> = 0.6, <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mi>g</mi> </msub> <mo>/</mo> <msub> <mi>l</mi> <mi>b</mi> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> </semantics></math>, <math display="inline"><semantics> <msub> <mi>M</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, and <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>t</mi> <mi>R</mi> </msubsup> </semantics></math> = 10).</p>
Full article ">Figure 10
<p>Graphs of <math display="inline"><semantics> <msub> <mi>W</mi> <mrow> <mi>N</mi> <mo>,</mo> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </semantics></math>-<math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>t</mi> <mi>R</mi> </msubsup> </semantics></math> for various MNO’s speeds: (<b>a</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, (<b>b</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.6, and (<b>c</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>N</mi> </msub> </semantics></math> = 0.9; ((∘) MNFA, <math display="inline"><semantics> <mfenced open="(" close=")"> <mo>▵</mo> </mfenced> </semantics></math> MNMA; (...) NRABT, and (—) NREBT; <math display="inline"><semantics> <mi>β</mi> </semantics></math> = 0.6, <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mi>g</mi> </msub> <mo>/</mo> <msub> <mi>l</mi> <mi>b</mi> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> </semantics></math>, <math display="inline"><semantics> <msub> <mi>M</mi> <mi>N</mi> </msub> </semantics></math> = 0.3, and <math display="inline"><semantics> <msubsup> <mover> <mi>k</mi> <mo>¯</mo> </mover> <mi>r</mi> <mi>R</mi> </msubsup> </semantics></math> = 10).</p>
Full article ">
22 pages, 6926 KiB  
Article
Vibration Characteristics Analysis of Immersed Tunnel Structures Based on a Viscoelastic Beam Model Embedded in a Fluid-Saturated Soil System Due to a Moving Load
by Hongyuan Huang, Yao Rong, Xiao Xiao and Bin Xu
Appl. Sci. 2023, 13(18), 10319; https://doi.org/10.3390/app131810319 - 14 Sep 2023
Viewed by 922
Abstract
This study aims to investigate the vibration responses on underwater immersed tunnels caused by moving loads, taking into account factors such as the viscoelastic characteristics of riverbed water, foundation soil, and the immersed tunnel itself. An ideal fluid medium is adopted to simulate [...] Read more.
This study aims to investigate the vibration responses on underwater immersed tunnels caused by moving loads, taking into account factors such as the viscoelastic characteristics of riverbed water, foundation soil, and the immersed tunnel itself. An ideal fluid medium is adopted to simulate the water, while a saturated porous medium is used to simulate the riverbed soil layer. The immersed tunnel structure is simplified as an infinitely long viscoelastic Euler beam, and the vibration effects are described by the theory of the standard linear solid model, taking into account structural damping. The coupled dynamic control equations were established by utilizing the displacement and stress conditions at the interface between the ideal fluid medium, the saturated porous medium, and the immersed tunnel structure. The equivalent stiffness of the riverbed water and site foundation was obtained. Furthermore, the numerical solutions of the tunnel displacement, internal forces, and pore pressure in the riverbed site were obtained in the time-space domain using the IFFT (Inverse Fast Fourier Transform) algorithm. The correctness of the model was validated by comparing the results with existing studies. The numerical results show that the riverbed water significantly reduces the Rayleigh wave velocity of the immersed tunnel structure in the riverbed-foundation system. Therefore, it is necessary to control the driving speed during high water levels. As the permeability of the saturated riverbed foundation increases, the vertical displacement, bending moment, and shear force of the beam in the immersed tunnel structure will increase. As the viscosity coefficient of the viscoelastic beam in the immersed tunnel structure increases, the vertical vibration amplitude of the beam will decrease, but further increasing the viscosity coefficient of the beam will have little effect on its vibration amplitude. Therefore, the standard solid model of the viscoelastic beam can effectively describe the creep and relaxation phenomena of materials and can objectively reflect the working conditions of the concrete structure of the immersed tunnel. Full article
(This article belongs to the Special Issue Urban Underground Engineering: Excavation, Monitoring, and Control)
Show Figures

Figure 1

Figure 1
<p>A schematic illustration of the water-soil-viscoelastic beam coupling system under a moving load.</p>
Full article ">Figure 2
<p>Schematic diagram of the immersed tunnel structure.</p>
Full article ">Figure 3
<p>Comparison with reference results. (<b>a</b>) horizontal displacement, (<b>b</b>) vertical displacement. The black square refers to Ref. [<a href="#B26-applsci-13-10319" class="html-bibr">26</a>].</p>
Full article ">Figure 4
<p>Vertical displacement change of the structural beam of the underwater immersed tunnel during different riverbed water depths. (<b>a</b>) <span class="html-italic">v</span><sub>c</sub> = 0.2 <span class="html-italic">v<sub>s</sub></span>, (<b>b</b>) <span class="html-italic">v<sub>c</sub></span> = 0.5 <span class="html-italic">v<sub>s</sub></span>, (<b>c</b>) <span class="html-italic">v<sub>c</sub></span> = 1.2 <span class="html-italic">v<sub>s</sub></span>.</p>
Full article ">Figure 4 Cont.
<p>Vertical displacement change of the structural beam of the underwater immersed tunnel during different riverbed water depths. (<b>a</b>) <span class="html-italic">v</span><sub>c</sub> = 0.2 <span class="html-italic">v<sub>s</sub></span>, (<b>b</b>) <span class="html-italic">v<sub>c</sub></span> = 0.5 <span class="html-italic">v<sub>s</sub></span>, (<b>c</b>) <span class="html-italic">v<sub>c</sub></span> = 1.2 <span class="html-italic">v<sub>s</sub></span>.</p>
Full article ">Figure 5
<p>Shear force change of the structural beam of the underwater immersed tunnel during different riverbed water depths. (<b>a</b>) v<sub>c</sub> = 0.2 <span class="html-italic">v<sub>s</sub></span>, (<b>b</b>) v<sub>c</sub> = 0.5 <span class="html-italic">v<sub>s</sub></span>, (<b>c</b>) v<sub>c</sub> = 1.2 <span class="html-italic">v<sub>s</sub></span>.</p>
Full article ">Figure 6
<p>Change of bending moment of the underwater immersed tunnel during different riverbed water depths. (<b>a</b>) v<sub>c</sub> = 0.2 <span class="html-italic">v<sub>s</sub></span>, (<b>b</b>) v<sub>c</sub> = 0.5 <span class="html-italic">v<sub>s</sub></span>, (<b>c</b>) v<sub>c</sub> = 1.2 <span class="html-italic">v<sub>s</sub></span>.</p>
Full article ">Figure 7
<p>Change of vertical displacement of riverbed foundation observation point during different riverbed water depths. (<b>a</b>) v<sub>c</sub> = 0.2 <span class="html-italic">v<sub>s</sub></span>, (<b>b</b>) v<sub>c</sub> = 0.5 <span class="html-italic">v<sub>s</sub></span>, (<b>c</b>) v<sub>c</sub> = 1.2 <span class="html-italic">v<sub>s</sub></span>.</p>
Full article ">Figure 7 Cont.
<p>Change of vertical displacement of riverbed foundation observation point during different riverbed water depths. (<b>a</b>) v<sub>c</sub> = 0.2 <span class="html-italic">v<sub>s</sub></span>, (<b>b</b>) v<sub>c</sub> = 0.5 <span class="html-italic">v<sub>s</sub></span>, (<b>c</b>) v<sub>c</sub> = 1.2 <span class="html-italic">v<sub>s</sub></span>.</p>
Full article ">Figure 8
<p>Horizontal displacement change of riverbed foundation observation point A (0.0 m, 2.0 m) at different riverbed water depths. (<b>a</b>) v<sub>c</sub> = 0.2 <span class="html-italic">v<sub>s</sub></span>, (<b>b</b>) v<sub>c</sub> = 0.5 <span class="html-italic">v<sub>s</sub></span>, (<b>c</b>) v<sub>c</sub> = 1.2 <span class="html-italic">v<sub>s</sub></span>.</p>
Full article ">Figure 9
<p>Change of hole pressure at the observation point of the riverbed foundation during different riverbed water depths. (<b>a</b>) v<sub>c</sub> = 0.2 <span class="html-italic">v<sub>s</sub></span>, (<b>b</b>) v<sub>c</sub> = 0.5 <span class="html-italic">v<sub>s</sub></span>, (<b>c</b>) v<sub>c</sub> = 1.2 <span class="html-italic">v<sub>s</sub></span>.</p>
Full article ">Figure 10
<p>Influence of different riverbed foundation permeability characteristics on the vibration of an underwater immersed tunnel structure beam. (<b>a</b>) vertical displacement, (<b>b</b>) shear force, (<b>c</b>) bending moment.</p>
Full article ">Figure 10 Cont.
<p>Influence of different riverbed foundation permeability characteristics on the vibration of an underwater immersed tunnel structure beam. (<b>a</b>) vertical displacement, (<b>b</b>) shear force, (<b>c</b>) bending moment.</p>
Full article ">Figure 11
<p>Influence of the permeability characteristics of different riverbed foundations on the vibration of the observation points in the foundation. (<b>a</b>) vertical displacement, (<b>b</b>) Horizontal displacement, (<b>c</b>) Hole pressure.</p>
Full article ">Figure 11 Cont.
<p>Influence of the permeability characteristics of different riverbed foundations on the vibration of the observation points in the foundation. (<b>a</b>) vertical displacement, (<b>b</b>) Horizontal displacement, (<b>c</b>) Hole pressure.</p>
Full article ">Figure 12
<p>Effect of the viscosity coefficient characteristics of different viscoelastic beams on the vibration of underwater immersed tunnel structure beams. (<b>a</b>) vertical displacement, (<b>b</b>) shear force, (<b>c</b>) bending moment.</p>
Full article ">Figure 12 Cont.
<p>Effect of the viscosity coefficient characteristics of different viscoelastic beams on the vibration of underwater immersed tunnel structure beams. (<b>a</b>) vertical displacement, (<b>b</b>) shear force, (<b>c</b>) bending moment.</p>
Full article ">Figure 13
<p>Effect of the viscosity coefficient characteristics of different viscoelastic beams on the vibration of the observation points in the foundation. (<b>a</b>) vertical displacement, (<b>b</b>) Horizontal displacement, (<b>c</b>) Hole pressure.</p>
Full article ">Figure 13 Cont.
<p>Effect of the viscosity coefficient characteristics of different viscoelastic beams on the vibration of the observation points in the foundation. (<b>a</b>) vertical displacement, (<b>b</b>) Horizontal displacement, (<b>c</b>) Hole pressure.</p>
Full article ">
17 pages, 2434 KiB  
Article
A Systematic Summary and Comparison of Scalar Diffraction Theories for Structured Light Beams
by Fuping Wu, Yi Luo and Zhiwei Cui
Photonics 2023, 10(9), 1041; https://doi.org/10.3390/photonics10091041 - 13 Sep 2023
Cited by 1 | Viewed by 1690
Abstract
Structured light beams have recently attracted enormous research interest for their unique properties and potential applications in optical communications, imaging, sensing, etc. Since most of these applications involve the propagation of structured light beams, which is accompanied by the phenomenon of diffraction, it [...] Read more.
Structured light beams have recently attracted enormous research interest for their unique properties and potential applications in optical communications, imaging, sensing, etc. Since most of these applications involve the propagation of structured light beams, which is accompanied by the phenomenon of diffraction, it is very necessary to employ diffraction theories to analyze the obstacle effects on structured light beams during propagation. The aim of this work is to provide a systematic summary and comparison of the scalar diffraction theories for structured light beams. We first present the scalar fields of typical structured light beams in the source plane, including the fundamental Gaussian beams, higher-order Hermite–Gaussian beams, Laguerre–Gaussian vortex beams, non-diffracting Bessel beams, and self-accelerating Airy beams. Then, we summarize and compare the main scalar diffraction theories of structured light beams, including the Fresnel diffraction integral, Collins formula, angular spectrum representation, and Rayleigh–Sommerfeld diffraction integral. Finally, based on these theories, we derive in detail the analytical propagation expressions of typical structured light beams under different conditions. In addition, the propagation of typical structured light beams is simulated. We hope this work can be helpful for the efficient study of the propagation of structured light beams. Full article
Show Figures

Figure 1

Figure 1
<p>Transverse intensity distributions of typical structured light beams under the paraxial approximation at different propagation distances. (<b>a1</b>–<b>a5</b>) <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>4</mn> <mi>λ</mi> </mrow> </semantics></math>, (<b>b1</b>–<b>b5</b>) <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>8</mn> <mi>λ</mi> </mrow> </semantics></math>, and (<b>c1</b>–<b>c5</b>) <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>12</mn> <mi>λ</mi> </mrow> </semantics></math>. Shown from left to right are the cases of the fundamental Gaussian beam, the Hermite–Gaussian beam, the Laguerre–Gaussian beam, the Bessel beam, and the Airy beam, respectively.</p>
Full article ">Figure 2
<p>Transverse intensity distributions of typical structured light beams beyond the paraxial approximation at different propagation distances. (<b>a1</b>–<b>a5</b>) <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>4</mn> <mi>λ</mi> </mrow> </semantics></math>, (<b>b1</b>–<b>b5</b>) <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>8</mn> <mi>λ</mi> </mrow> </semantics></math>, and (<b>c1</b>–<b>c5</b>) <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>12</mn> <mi>λ</mi> </mrow> </semantics></math>. Shown from left to right are the cases of the fundamental Gaussian beam, the Hermite–Gaussian beam, the Laguerre–Gaussian beam, the Bessel beam, and the Airy beam, respectively.</p>
Full article ">Figure 3
<p>Illustrations of the propagation of typical structured light beams in a gradient-index medium. (<b>a</b>) Fundamental Gaussian beam, (<b>b</b>) Hermite–Gaussian beam, (<b>c</b>) Laguerre–Gaussian beam, (<b>d</b>) Bessel beam, and (<b>e</b>) Airy beam.</p>
Full article ">
Back to TopTop