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

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (845)

Search Parameters:
Keywords = direct shear test

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 12122 KiB  
Article
Microstructural and Mechanical Characterization of Nb-Doped MoS2 Coatings Deposited on H13 Tool Steel Using Nb-Based Interlayers
by Miguel R. Danelon, Newton K. Fukumasu, Angelo A. Carvalho, Ronnie R. Rego, Izabel F. Machado, Roberto M. Souza and André P. Tschiptschin
Coatings 2025, 15(1), 57; https://doi.org/10.3390/coatings15010057 - 6 Jan 2025
Viewed by 305
Abstract
Molybdenum disulfide is a 2D material with excellent lubricant properties, resulting from weak van der Waals forces between lattice layers and shear-induced crystal orientation. The low forces needed to shear the MoS2 crystal layers grant the tribological system low coefficients of friction [...] Read more.
Molybdenum disulfide is a 2D material with excellent lubricant properties, resulting from weak van der Waals forces between lattice layers and shear-induced crystal orientation. The low forces needed to shear the MoS2 crystal layers grant the tribological system low coefficients of friction (COF). However, film oxidation harms its efficacy in humid atmospheres, leading to an increased COF and poor surface adhesion, making its use preferable in dry or vacuum conditions. To overcome these challenges, doping MoS2 with elements such as Nb, Ti, C, and N emerges as a promising solution. Nevertheless, the adhesion of these coatings to a steel substrate presents challenges and strategies involving the reduction in residual stresses and increased chemical affinity to the substrate by using niobium-based materials as interlayers. In this study, Nb-doped MoS2 films were deposited on H13 steel and silicon wafers using the pulsed direct current balanced magnetron sputtering technique. Different niobium-based interlayers (pure Nb and NbN) were deposited to evaluate the adhesion properties of Nb-doped MoS2 coatings. Unlubricated scratch tests, conducted at room temperature and relative humidity under a progressive load, were performed to analyze the COF and adhesion of the coating. Instrumented indentation tests were conducted to assess the hardness and elastic modulus of the coatings. The microstructure of the coatings was obtained by Scanning Electron Microscopy (SEM), Scanning Transmission Electron Microscopy (STEM), and Transmission Electron Microscopy (TEM), with Energy-Dispersive X-Ray Spectroscopy (EDS). Results indicated that niobium doping on MoS2 coatings changes the structure from crystalline to amorphous. Additionally, the Nb concentration of the Nb:MoS2 coating changed the mechanical properties, leading to different cohesive failures by different loads during the scratch tests. Results have also indicated that an NbN interlayer optimally promoted the adhesion of the film. This result is justified by the increase in hardness led by higher Nb concentrations, enhancing the load-bearing capacity of the coating. It is concluded that niobium-based materials can be used to enhance the adhesion properties of Nb-doped MoS2 films and improve their tribological performance. Full article
(This article belongs to the Special Issue Friction, Wear, Lubrication and Mechanics of Surfaces and Interfaces)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Scheme of deposition and (<b>B</b>) architecture of Nb:MoS<sub>2</sub> coatings.</p>
Full article ">Figure 2
<p>Microstructures of Nb:MoS<sub>2</sub> coatings deposited under different powers applied to the Nb target: (<b>A</b>) 25 W, (<b>B</b>) 30 W, (<b>C</b>) 35 W, and (<b>D</b>) 40 W.</p>
Full article ">Figure 3
<p>Microstructure of Nb25NbN300 sample at SEM.</p>
Full article ">Figure 4
<p>FEG Nb30NbN300 film microstructure. The Pt light gray layer was deposited on the surface to protect the film from being damaged by the focused ion beam.</p>
Full article ">Figure 5
<p>EDS analysis of the cross-section of different layers.</p>
Full article ">Figure 6
<p>TEM image of the transition between Pt, MoS<sub>2</sub>, and Nb:MoS<sub>2</sub> layers.</p>
Full article ">Figure 7
<p>TEM images of the (<b>A</b>) pure MoS<sub>2</sub> layer, (<b>B</b>) Nb:MoS<sub>2</sub> film, and (<b>C</b>) scheme of the crystal structure of a MoS<sub>2</sub> monolayer showing a layer of molybdenum (blue) sandwiched between two layers of sulfur (yellow).</p>
Full article ">Figure 8
<p>TEM images of the (<b>A</b>) MoS<sub>2</sub> and (<b>C</b>) Nb:MoS<sub>2</sub> layers. Red circles indicate regions for SAED analysis, with corresponding patterns in (<b>B</b>) and (<b>D</b>), respectively.</p>
Full article ">Figure 8 Cont.
<p>TEM images of the (<b>A</b>) MoS<sub>2</sub> and (<b>C</b>) Nb:MoS<sub>2</sub> layers. Red circles indicate regions for SAED analysis, with corresponding patterns in (<b>B</b>) and (<b>D</b>), respectively.</p>
Full article ">Figure 9
<p>The transition between the NbN and Nb layers shows partial epitaxial growth. The blue line was included as an indication of the interface between both layers.</p>
Full article ">Figure 10
<p>TEM images of the (<b>A</b>) NbN and (<b>B</b>) Nb layers. Red circles indicate regions for SAED analysis, with corresponding patterns in (<b>C</b>) and (<b>D</b>), respectively.</p>
Full article ">Figure 11
<p>Raman spectroscopy of niobium nitride from NbN300 sample. The insert was taken from [<a href="#B58-coatings-15-00057" class="html-bibr">58</a>].</p>
Full article ">Figure 12
<p>Scratch images of samples (<b>A</b>) Nb25, (<b>B</b>) Nb30, (<b>C</b>) Nb35, (<b>D</b>) Nb40, and (<b>E</b>) Nb25NbN300.</p>
Full article ">Figure 13
<p>Graphs of COF, tangential force, and normal force as a function of the distance of the scratch for the samples (<b>A</b>) Nb25, (<b>B</b>) Nb30, (<b>C</b>) Nb35, (<b>D</b>) Nb40, and (<b>E</b>) Nb25NbN300.</p>
Full article ">Figure 14
<p>Characterization of cohesive failure of coatings by (<b>A</b>) optical microscope and (<b>B</b>) SEM.</p>
Full article ">Figure 15
<p>Characterization of adhesive failure by (<b>A</b>) optical microscope, (<b>B</b>) SEM and EDS, and (<b>C</b>) combined analysis of COF curve and scratch image.</p>
Full article ">Figure 15 Cont.
<p>Characterization of adhesive failure by (<b>A</b>) optical microscope, (<b>B</b>) SEM and EDS, and (<b>C</b>) combined analysis of COF curve and scratch image.</p>
Full article ">
27 pages, 10001 KiB  
Article
Influential Mechanisms of Roughness on the Cyclic Shearing Behavior of the Interfaces Between Crushed Mudstone and Steel-Cased Rock-Socketed Piles
by Yue Liang, Jianlu Zhang, Bin Xu, Zeyu Liu, Lei Dai and Kui Wang
Buildings 2025, 15(1), 141; https://doi.org/10.3390/buildings15010141 - 5 Jan 2025
Viewed by 465
Abstract
In the waterway construction projects of the upper reaches of the Yangtze River, crushed mudstone particles are widely used to backfill the foundations of rock-socketed concrete-filled steel tube (RSCFST) piles, a structure widely adopted in port constructions. In these projects, the steel–mudstone interfaces [...] Read more.
In the waterway construction projects of the upper reaches of the Yangtze River, crushed mudstone particles are widely used to backfill the foundations of rock-socketed concrete-filled steel tube (RSCFST) piles, a structure widely adopted in port constructions. In these projects, the steel–mudstone interfaces experience complex loading conditions, and the surface profile tends to vary within certain ranges during construction and operation. The changes in boundary conditions and material profile significantly impact the bearing performance of these piles when subjected to cyclic loads, such as ship impacts, water level fluctuations, and wave-induced loads. Therefore, it is necessary to investigate the shear characteristics of the RSCFST pile–soil interface under cyclic vertical loading, particularly in relation to varying deformation levels in the steel casing’s outer profile. In this study, a series of cyclic direct shear tests are carried out to investigate the influential mechanisms of roughness on the cyclic behavior of RSCFST pile–soil interfaces. The impacts of roughness on shear stress, shear stiffness, damping ratio, normal stress, and particle breakage ratio are discussed separately and can be summarized as follows: (1) During the initial phase of cyclic shearing, increased roughness correlates with higher interfacial shear strength and anisotropy, but also exacerbates interfacial particle breakage. Consequently, the sample undergoes more significant shear contraction, leading to reduced interfacial shear strength and anisotropy in the later stages. (2) The damping ratio of the rough interface exhibits an initial increase followed by a decrease, while the smooth interface demonstrates the exact opposite trend. The variation in damping ratio characteristics corresponds to the transition from soil–structure to soil–soil interfacial shearing. (3) Shear contraction is more pronounced in rough interface samples compared to the smooth interface, indicating that particle breakage has a greater impact on soil shear contraction compared to densification. Full article
(This article belongs to the Special Issue Structural Mechanics Analysis of Soil-Structure Interaction)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of RSCFST piles.</p>
Full article ">Figure 2
<p>Schematic diagram of the CNS boundary condition.</p>
Full article ">Figure 3
<p>Mudstone particles used in the tests. (<b>a</b>) Particle size distribution curve of soil specimens; (<b>b</b>) physical graph of mudstone particles.</p>
Full article ">Figure 4
<p>Surface profile of steel plate.</p>
Full article ">Figure 5
<p>Large-scale CNS cyclic direct shear apparatus. (<b>a</b>) Schematic diagram; (<b>b</b>) physical graph.</p>
Full article ">Figure 6
<p>Schematic diagram of the cyclic shear path.</p>
Full article ">Figure 7
<p>Shear stress versus shear displacement. (<b>a</b>–<b>d</b>) INS = 300 kPa; (<b>e</b>–<b>h</b>) INS = 400 kPa; (<b>i</b>–<b>l</b>) INS = 500 kPa.</p>
Full article ">Figure 7 Cont.
<p>Shear stress versus shear displacement. (<b>a</b>–<b>d</b>) INS = 300 kPa; (<b>e</b>–<b>h</b>) INS = 400 kPa; (<b>i</b>–<b>l</b>) INS = 500 kPa.</p>
Full article ">Figure 7 Cont.
<p>Shear stress versus shear displacement. (<b>a</b>–<b>d</b>) INS = 300 kPa; (<b>e</b>–<b>h</b>) INS = 400 kPa; (<b>i</b>–<b>l</b>) INS = 500 kPa.</p>
Full article ">Figure 8
<p>Maximum shear stress versus number of cycles in the positive and negative direction. (<b>a</b>–<b>d</b>) INS = 300 kPa; (<b>e</b>–<b>h</b>) INS = 400 kPa; (<b>i</b>–<b>l</b>) INS = 500 kPa.</p>
Full article ">Figure 8 Cont.
<p>Maximum shear stress versus number of cycles in the positive and negative direction. (<b>a</b>–<b>d</b>) INS = 300 kPa; (<b>e</b>–<b>h</b>) INS = 400 kPa; (<b>i</b>–<b>l</b>) INS = 500 kPa.</p>
Full article ">Figure 9
<p>Schematic diagram of shear stiffness and damping ratio.</p>
Full article ">Figure 10
<p>Shear stiffness versus number of cycles. (<b>a</b>) INS = 300 kPa; (<b>b</b>) INS = 400 kPa; (<b>c</b>) INS = 500 kPa.</p>
Full article ">Figure 11
<p>Damping ratio versus number of cycles. (<b>a</b>) INS = 300 kPa; (<b>b</b>) INS = 400 kPa; (<b>c</b>) INS = 500 kPa.</p>
Full article ">Figure 12
<p>Normal stress versus shear displacement. (<b>a–d</b>) INS = 300 kPa; (<b>e–h</b>) INS = 400 kPa; (<b>i–l</b>) INS = 500 kPa.</p>
Full article ">Figure 12 Cont.
<p>Normal stress versus shear displacement. (<b>a–d</b>) INS = 300 kPa; (<b>e–h</b>) INS = 400 kPa; (<b>i–l</b>) INS = 500 kPa.</p>
Full article ">Figure 12 Cont.
<p>Normal stress versus shear displacement. (<b>a–d</b>) INS = 300 kPa; (<b>e–h</b>) INS = 400 kPa; (<b>i–l</b>) INS = 500 kPa.</p>
Full article ">Figure 13
<p>Normal stress versus number of cycles. (<b>a</b>) INS = 300 kPa; (<b>b</b>) INS = 400 kPa; (<b>c</b>) INS = 500 kPa.</p>
Full article ">Figure 14
<p>Normal stress attenuation ratio versus roughness.</p>
Full article ">Figure 15
<p>Particle size distribution curve before and after shear test. (<b>a</b>) INS = 300 kPa; (<b>b</b>) INS = 400 kPa; (<b>c</b>) INS = 500 kPa.</p>
Full article ">Figure 16
<p>Particle breakage ratio versus roughness. (<b>a</b>) INS = 300 kPa; (<b>b</b>) INS = 400 kPa; (<b>c</b>) INS = 500 kPa.</p>
Full article ">Figure 17
<p>Interface friction angle versus number of cycles.</p>
Full article ">Figure 18
<p>Contact type of the shear zone before and after test. (<b>a</b>) Laboratory graph; (<b>b</b>) schematic diagram.</p>
Full article ">
14 pages, 3232 KiB  
Article
Experimental Investigation on Unloading-Induced Sliding Behavior of Dry Sands Subjected to Constant Shear Force
by Wengang Dang, Kang Tao, Jinyang Fu and Bangbiao Wu
Appl. Sci. 2025, 15(1), 401; https://doi.org/10.3390/app15010401 - 3 Jan 2025
Viewed by 394
Abstract
Infilled joints or faults are often subjected to long-term stable shear forces, and nature surface processes of normal unloading can change the frictional balance. Therefore, it is essential to study the sliding behavior of such granular materials under such unloading conditions, since they [...] Read more.
Infilled joints or faults are often subjected to long-term stable shear forces, and nature surface processes of normal unloading can change the frictional balance. Therefore, it is essential to study the sliding behavior of such granular materials under such unloading conditions, since they are usually the filling matter. We conducted two groups of normal unloading direct shear tests considering two variables: unloading rate and the magnitude of constant shear force. Dry sands may slide discontinuously during normal unloading, and the slip velocity does not increase uniformly with unloading time. Due to horizontal particle interlacing and normal relaxation, there will be sliding velocity fluctuations and even temporary intermissions. At the stage of sliding acceleration, the normal force decreases with a higher unloading rate and increases with a larger shear force at the same sliding velocity. The normal forces obtained from the tests are less than those calculated by Coulomb’s theory in the conventional constant-rate shear test. Under the same unloading rate, the range of apparent friction coefficient variation is narrower under larger shear forces. This study has revealed the movement patterns of natural granular layers and is of enlightening significance in the prevention of corresponding geohazards. Full article
(This article belongs to the Topic Geotechnics for Hazard Mitigation)
Show Figures

Figure 1

Figure 1
<p>Testing apparatus and experimental materials: (<b>a</b>) the DJZ-500 shear box device; (<b>b</b>) dry sands used in tests; and (<b>c</b>) grain grading curve of the sand samples.</p>
Full article ">Figure 2
<p>Experimental configuration: (<b>a</b>) setup of direct shear test under unloading normal force and constant shear force; (<b>b</b>) normal and shear force application scheme during three-stage loading process.</p>
Full article ">Figure 3
<p>Normal force variation as function of elapsed time and experimental results of normal and sliding displacement versus time for each test. (<b>a</b>) Normal force, normal displacement, and sliding displacement versus time for different unloading rates (Group A). (<b>b</b>) Normal force, normal displacement, and sliding displacement versus time for different shear force (Group B).</p>
Full article ">Figure 4
<p>Sliding velocity versus time for each test since the beginning moment of normal unloading. (<b>a</b>) Sliding velocity versus time for different unloading rates. (<b>b</b>) Sliding velocity versus time for different shear force.</p>
Full article ">Figure 5
<p>(<b>a</b>) Relationship between shear displacement and shear force under <span class="html-italic">F</span><sub>N</sub> = 30 kN in a conventional direct shear test. Within the displacement range, the shear strength of the granular material keeps increasing with larger shear displacement. (<b>b</b>) The peak shear force (shear strength) for different normal force in 0.0833 mm/s shear-displacement-controlled conventional direct shear test. (<b>c</b>) Shear displacement (solid lines) and normal displacement (dash lines) versus time for low (0.08 kN/s) and high (0.8 kN/s) unloading rates extracted from <a href="#applsci-15-00401-f003" class="html-fig">Figure 3</a>a and illustration to explain sliding deceleration/intermission.</p>
Full article ">Figure 6
<p>Variation in normal force at different sliding velocities. (<b>a</b>) For different unloading rate (Group A) and (<b>b</b>) for different shear force (Group B).</p>
Full article ">Figure 7
<p>Normal force versus unloading rate and shear force at sliding velocity of (<b>a</b>) 3.7 mm/s, (<b>b</b>) 2.5 mm/s, and (<b>c</b>) 1.5 mm/s.</p>
Full article ">Figure 8
<p>(<b>a</b>) Apparent friction coefficient when the sliding velocity reaches 3.7 mm/s (i.e., ultimate value in the test) for different unloading rates and different shear force. (<b>b</b>) Apparent friction coefficient versus unloading rate and shear force at sliding velocity of 3.7 mm/s. (<b>c</b>) The variation in friction coefficient versus normal unloading time for different shear force (the recorded data in Group B). The different colors represent different single tests.</p>
Full article ">
22 pages, 5922 KiB  
Article
Predictive Modeling and Experimental Analysis of Cyclic Shear Behavior in Sand–Fly Ash Mixtures
by Özgür Yıldız and Ali Fırat Çabalar
Appl. Sci. 2025, 15(1), 353; https://doi.org/10.3390/app15010353 - 2 Jan 2025
Viewed by 288
Abstract
This study presents a comprehensive investigation into the cyclic shear behavior of sand–fly ash mixtures through experimental and data-driven modeling approaches. Cyclic direct shear tests were conducted on mixtures containing fly ash at 0%, 2.5%, 5%, 10%, 15%, and 20% by weight to [...] Read more.
This study presents a comprehensive investigation into the cyclic shear behavior of sand–fly ash mixtures through experimental and data-driven modeling approaches. Cyclic direct shear tests were conducted on mixtures containing fly ash at 0%, 2.5%, 5%, 10%, 15%, and 20% by weight to examine the influence of fly ash content on the shear behavior under cyclic loading conditions. The tests were carried out under a constant stress of 100 kPa to simulate field-relevant stress conditions. Results revealed that the fly ash content initially reduces shear strength at lower additive contents, but shear strength increases and reaches a maximum at 20% fly ash content. The findings highlight the trade-offs in mechanical behavior associated with varying fly ash proportions. To enhance the understanding of cyclic shear behavior, a Nonlinear Autoregressive Model with External Input (NARX) model was employed. Using data from the loading cycles as input, the NARX model was trained to predict the final shear response under cyclic conditions. The model demonstrated exceptional predictive performance, achieving a coefficient of determination (R2) of 0.99, showcasing its robustness in forecasting the cyclic shear performance based on the composition of the mixtures. The insights derived from this research underscore the potential of incorporating fly ash in sand mixtures for soil stabilization in geotechnical engineering. Furthermore, the integration of advanced machine learning techniques such as NARX models offers a powerful tool for predicting the behavior of soil mixtures, facilitating more effective and data-driven decision-making in geotechnical applications. Evidently, this study not only advances the understanding of cyclic shear behavior in fly ash–sand mixtures but also provides a framework for employing data-driven methodologies to address complex geotechnical challenges. Full article
Show Figures

Figure 1

Figure 1
<p>Grain size distribution of Trakya sand (TS).</p>
Full article ">Figure 2
<p>SEM pictures of Trakya sand grains.</p>
Full article ">Figure 3
<p>Materials and equipment used: (<b>a</b>) fly ash, (<b>b</b>) Trakya sand, (<b>c</b>) cyclic direct shear testing equipment, (<b>d</b>) preparing specimens inside the shear box, (<b>e</b>) specimens after test.</p>
Full article ">Figure 4
<p>The variation in the volumetric strain at each cycle of loading.</p>
Full article ">Figure 5
<p>The variation in the volumetric strain of the tested specimens of; (<b>a</b>) Clean sand and mixtures of (<b>b</b>) 2.5 (%) FA, (<b>c</b>) 5 (%) FA, (<b>d</b>) 10 (%) FA, (<b>e</b>) 15 (%) FA and (<b>f</b>) 20 (%) FA.</p>
Full article ">Figure 6
<p>First-quarter cyclic shear behavior of tested specimens.</p>
Full article ">Figure 7
<p>Variations in shear stress with horizontal strain for; (<b>a</b>) Clean sand and mixtures of (<b>b</b>) 2.5 (%) FA, (<b>c</b>) 5 (%) FA, (<b>d</b>) 10 (%) FA, (<b>e</b>) 15 (%) FA and (<b>f</b>) 20 (%) FA.</p>
Full article ">Figure 7 Cont.
<p>Variations in shear stress with horizontal strain for; (<b>a</b>) Clean sand and mixtures of (<b>b</b>) 2.5 (%) FA, (<b>c</b>) 5 (%) FA, (<b>d</b>) 10 (%) FA, (<b>e</b>) 15 (%) FA and (<b>f</b>) 20 (%) FA.</p>
Full article ">Figure 8
<p>Schematical demonstration of discretization of hysteresis loop.</p>
Full article ">Figure 9
<p>The flowchart of the analysis.</p>
Full article ">Figure 10
<p>The architecture of the NARX model.</p>
Full article ">Figure 11
<p>Best training performance of NARX model.</p>
Full article ">Figure 12
<p>Training state of NARX model.</p>
Full article ">Figure 13
<p>Response plot of NARX model.</p>
Full article ">Figure 14
<p>Predicted and measured stress–strain responses for specimens during fifth cycle; (<b>a</b>) Clean sand and mixtures of (<b>b</b>) 2.5 (%) FA, (<b>c</b>) 5 (%) FA, (<b>d</b>) 10 (%) FA, (<b>e</b>) 15 (%) FA and (<b>f</b>) 20 (%) FA.</p>
Full article ">Figure 14 Cont.
<p>Predicted and measured stress–strain responses for specimens during fifth cycle; (<b>a</b>) Clean sand and mixtures of (<b>b</b>) 2.5 (%) FA, (<b>c</b>) 5 (%) FA, (<b>d</b>) 10 (%) FA, (<b>e</b>) 15 (%) FA and (<b>f</b>) 20 (%) FA.</p>
Full article ">
23 pages, 3901 KiB  
Article
Hypoplastic Modeling of Soil–Structure Contact Surface Considering Initial Anisotropy and Roughness
by Jingtao Yu, Junwang Cao, Zixuan Chen, Jintao Zhu, Yulong Zhang and Pengqiang Yu
Appl. Sci. 2025, 15(1), 244; https://doi.org/10.3390/app15010244 - 30 Dec 2024
Viewed by 285
Abstract
The development of a constitutive model for soil–structure contact surfaces remains a pivotal area of research within the field of soil–structure interaction. Drawing from the Gudehus–Bauer sand hypoplasticity model, this paper employs a technique that reduces the stress tensor and strain rate tensor [...] Read more.
The development of a constitutive model for soil–structure contact surfaces remains a pivotal area of research within the field of soil–structure interaction. Drawing from the Gudehus–Bauer sand hypoplasticity model, this paper employs a technique that reduces the stress tensor and strain rate tensor components to formulate a hypoplastic model tailored for sand–structure interfaces. To capture the influence of initial anisotropy, a deposition direction peak stress coefficient is incorporated; meanwhile, a friction parameter is introduced to address the surface roughness of the contact. Consequently, a comprehensive hypoplastic constitutive model is developed that takes into account both initial anisotropy and roughness. Comparative analysis with experimental data from soils on contact surfaces with diverse boundary conditions and levels of roughness indicates that the proposed model accurately forecasts shear test outcomes across various contact surfaces. Utilizing the finite element software ABAQUS 2021, an FRIC subroutine was developed, which, through simulating direct shear tests on sand–structure contact surfaces, has proven its efficacy in predicting the shear behavior of these interfaces. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the contact surface coordinate system.</p>
Full article ">Figure 2
<p>Schematic diagram of contact surface stress and strain components.</p>
Full article ">Figure 3
<p>Variations in peak stress coefficient with the deposition direction.</p>
Full article ">Figure 4
<p>Diagram of peak-to-valley distance of the structural surface.</p>
Full article ">Figure 5
<p>Schematic diagram of strain–displacement relationship at the contact surface.</p>
Full article ">Figure 6
<p>Comparison of experimental results with simulation results for the contact surface between Toyoura sand and low-carbon steel plates.</p>
Full article ">Figure 7
<p>Experimental results and simulation results for Changping composite gravel–artificial rough steel plate contact surface under different roughness conditions.</p>
Full article ">Figure 8
<p>Experimental results and simulation results for Changping uniform gravel–artificial rough steel plate contact surface under different roughness conditions.</p>
Full article ">Figure 9
<p>Experimental results and simulation results for uniform density sand–concrete contact surface under staged stress path.</p>
Full article ">Figure 10
<p>Basic algorithm flow of the finite element method for the hypoplasticity model of contact surfaces.</p>
Full article ">Figure 11
<p>Schematic diagram of the geometric structure and boundary conditions of the finite element model for direct shear test: (<b>a</b>) front view and (<b>b</b>) side view.</p>
Full article ">Figure 12
<p>Model grid schematic diagram.</p>
Full article ">Figure 13
<p>Comparison between the numerical results of the hypoplastic model for contact surfaces and the theoretical solution.</p>
Full article ">Figure 14
<p>Shear stress–displacement relationship curves for contact surfaces under different interface roughness conditions: (<b>a</b>) e = 0.65 and (<b>b</b>) e = 0.95.</p>
Full article ">Figure 15
<p>Shear stress–displacement relationship curves for contact surfaces under different initial deposition angle conditions: (<b>a</b>) e = 0.65 and (<b>b</b>) e = 0.90.</p>
Full article ">Figure 16
<p>Comparison of shear stress–displacement relations between the Mohr–Coulomb model and the hypoplasticity model.</p>
Full article ">
21 pages, 7242 KiB  
Article
Innovative Fly-Ash-Based Soil Crust Rehabilitation: Enhancing Wind Erosion Resistance in Gravel-Layered Desert Mining Areas
by Yu Zheng, Weiming Guan, Jingwen Li, Zhenqi Hu, Gensheng Li, Meng Xie and Xuewei Zhang
Land 2025, 14(1), 36; https://doi.org/10.3390/land14010036 - 27 Dec 2024
Viewed by 460
Abstract
Gravel layers are vital ecological barriers in Gobi Desert mining areas. However, open-pit activities increase wind and soil erosion. Thus, the effects of fly ash addition, water content, and compaction on the shear strength and wind erosion resistance of soil crusts were explored [...] Read more.
Gravel layers are vital ecological barriers in Gobi Desert mining areas. However, open-pit activities increase wind and soil erosion. Thus, the effects of fly ash addition, water content, and compaction on the shear strength and wind erosion resistance of soil crusts were explored by compaction tests, direct shear tests, and wind tunnel experiments. (1) The results of the direct shear test and vane shear test show that the modified soil sample achieved the maximum shear strength under the conditions of 15% fly ash content, 13% water content, and 3 compaction cycles. (2) The results of the wind tunnel test indicate that the wind erosion resistance of the gravel layer soil crust was improved after fly ash treatment. Compared to the untreated soil crust, the wind erosion amount of the treated soil was reduced by 23%. (3) Microscopic analysis revealed that hydration products from fly ash filled the soil pores, enhancing particle bonding and soil structure, using a scanning electron microscope (SEM) and an X-ray fluorescence spectrometer (XRF). (4) Considering the water scarcity in the Eastern Junggar Coalfield of China, a revised rehabilitation scheme was selected, involving 11% water content and single compaction, offering a balance between performance and economic efficiency. This study provides a novel approach to gravel layer restoration in arid mining regions using fly ash as a soil stabilizer, offering a sustainable method to enhance wind erosion resistance and promote fly ash recycling. Full article
Show Figures

Figure 1

Figure 1
<p>Locations of the sampling sites. (<b>a</b>) China. (<b>b</b>) Xinjiang Uighur Autonomous Region. (<b>c</b>) Eastern Junggar Coalfield.</p>
Full article ">Figure 2
<p>Real view of the study area. (<b>a</b>) Aerial drone view of the study area. (<b>b</b>) Local surface map. (<b>c</b>) Section view.</p>
Full article ">Figure 3
<p>Experimental process flow diagram.</p>
Full article ">Figure 4
<p>Homemade compaction device. (<b>a</b>) Actual image. (<b>b</b>) Schematic diagram.</p>
Full article ">Figure 5
<p>Mini vane shear apparatus.</p>
Full article ">Figure 6
<p>Miniature wind tunnel test apparatus. (<b>a</b>) Actual image. (<b>b</b>) Schematic diagram.</p>
Full article ">Figure 7
<p>Compaction curves of soil samples with varying fly ash proportions.</p>
Full article ">Figure 8
<p>Relationship curves among optimal moisture content, maximum dry density, and fly ash proportion.</p>
Full article ">Figure 9
<p>Fly ash proportion and shear strength curve.</p>
Full article ">Figure 10
<p>The relationship curve between the shear strength of soil with different fly ash contents and vertical pressure.</p>
Full article ">Figure 11
<p>The relationship curves between cohesion, internal friction angle, and fly ash content.</p>
Full article ">Figure 12
<p>Water content and shear strength curve.</p>
Full article ">Figure 13
<p>The relationship curve between the shear strength of soil at different water contents and vertical pressure.</p>
Full article ">Figure 14
<p>The relationship curves between cohesion, internal friction angle, and water content.</p>
Full article ">Figure 15
<p>The relationship curves between shear strength, bulk density, and the number of compaction cycles.</p>
Full article ">Figure 16
<p>Comparison of the material composition of gravel-layered soil before and after reconstruction. (<b>a</b>) The natural gravel-layered soil. (<b>b</b>) The reconstituted gravel-layered soil.</p>
Full article ">Figure 17
<p>Microstructural images of soil sample with 0% fly ash content. (<b>a</b>) 500×. (<b>b</b>) 1000×.</p>
Full article ">Figure 18
<p>Microstructural images of soil sample with 15% fly ash content. (<b>a</b>) 500×. (<b>b</b>) 1000×.</p>
Full article ">Figure 19
<p>Cumulative wind erosion relationship diagram.</p>
Full article ">
14 pages, 3806 KiB  
Article
Study on the Influence of Water Content on the Shear Behavior of the Soil–Structure Interface Under a Temperature Field
by Jian Chen, Hao Jiang, Yongde Liu, Yanting Wu, Xuan Zhang and Weidong Pan
Buildings 2025, 15(1), 1; https://doi.org/10.3390/buildings15010001 - 24 Dec 2024
Viewed by 363
Abstract
Energy piles are highly favored for their excellent, low energy consumption in providing heating for public residences. The temperature field changes the activity of the diffuse double electric layer (DEL) on the particle surface, thereby altering the distribution of the stress field in [...] Read more.
Energy piles are highly favored for their excellent, low energy consumption in providing heating for public residences. The temperature field changes the activity of the diffuse double electric layer (DEL) on the particle surface, thereby altering the distribution of the stress field in the soil and ultimately affecting the mechanical properties of the interface between the energy pile and the soil. Therefore, studying the influence of water content on the mechanical behavior of the soil–structure interface in the temperature field is crucial for energy pile safety. This study used a modified temperature-controlled direct shear apparatus to obtain the influence of water content and temperature on the shear behavior of the soil–structure interface. Then, the test results were analyzed and discussed. Finally, three results were obtained: (1) The water content of bentonite (wbent) had a significant impact on the shear stress–shear displacement curve of the soil–structure interface; when the wbent was less than the wp of the bentonite, the τ-l curve exhibited a softening response, then displayed a hardening response. (2) The shear strength of the soil–structure interface gradually decreased with the increase of wbent. (3) The shear strength of the soil–structure interface increased with increasing temperature under various wbent and vertical loads. Full article
Show Figures

Figure 1

Figure 1
<p>Particle size distribution curve of the sand.</p>
Full article ">Figure 2
<p>Self-improved temperature-controlled direct shear apparatus: (<b>a</b>) instrument schematic diagram; (<b>b</b>) physical image.</p>
Full article ">Figure 3
<p>The <span class="html-italic">τ</span>-<span class="html-italic">l</span> relationship curve of different temperatures at different moisture contents: (<b>a</b>) w<sub>bent</sub> = 40%; (<b>b</b>) <span class="html-italic">w</span><sub>bent</sub> = 55%; (<b>c</b>) <span class="html-italic">w</span><sub>bent</sub> = 70%; (<b>d</b>) <span class="html-italic">w</span><sub>bent</sub> = 85%; (<b>e</b>) <span class="html-italic">w</span><sub>bent</sub> = 100%.</p>
Full article ">Figure 4
<p><span class="html-italic">τ</span><sub>max</sub>–<span class="html-italic">w</span><sub>bent</sub> curve of soil–structure interface: (<b>a</b>) P = 12.5 kPa; (<b>b</b>) P = 25 kPa; (<b>c</b>) P = 37.5 kPa; (<b>d</b>) P = 50 kPa.</p>
Full article ">Figure 5
<p>Shear envelope of soil–structure interface (<span class="html-italic">w</span><sub>bent</sub> = 100%).</p>
Full article ">Figure 6
<p>Relationship curve of <span class="html-italic">c</span><sub>i</sub>–<span class="html-italic">w</span><sub>bent</sub> of the soil–structure interface (CC = clay content).</p>
Full article ">Figure 7
<p>Relationship curve of <span class="html-italic">φ</span><sub>i</sub>–<span class="html-italic">w</span><sub>bent</sub> of soil–structure interface (R = roughness; T = temperature).</p>
Full article ">Figure 8
<p>Schematic diagram of the impact of <span class="html-italic">w</span><sub>bent</sub> changes on interfaces.</p>
Full article ">Figure 9
<p>The influence of temperature changes on the DEL.</p>
Full article ">
15 pages, 6531 KiB  
Article
Preparation and Performance Study of Novel Foam Vegetation Concrete
by Teng Zhang, Tianbin Li, Hua Xu, Mengyun Wang and Lingling Lu
Materials 2024, 17(24), 6295; https://doi.org/10.3390/ma17246295 - 23 Dec 2024
Viewed by 463
Abstract
Vegetation concrete is one of the most widely used substrates in ecological slope protection, but its practical application often limits the growth and nutrient uptake of plant roots due to consolidation problems, which affects the effectiveness of slope protection. This paper proposed the [...] Read more.
Vegetation concrete is one of the most widely used substrates in ecological slope protection, but its practical application often limits the growth and nutrient uptake of plant roots due to consolidation problems, which affects the effectiveness of slope protection. This paper proposed the use of a plant protein foaming agent as a porous modifier to create a porous, lightweight treatment for vegetation concrete. Physical performance tests, direct shear tests, plant growth tests, and scanning electron microscopy experiments were conducted to compare and analyze the physical, mechanical, microscopic characteristics, and phyto-capabilities of differently treated vegetation concrete. The results showed that the higher the foam content, the more significant the porous and lightweight properties of the vegetation concrete. When the foam volume was 50%, the porosity increased by 106.05% compared to the untreated sample, while the volume weight decreased by 20.53%. The shear strength, cohesion, and internal friction angle of vegetation concrete all showed a decreasing trend with increasing foaming agent content. Festuca arundinacea grew best under the 30% foaming agent treatment, with germinative energy, germinative percentage, plant height, root length, and underground biomass increasing by 6.31%, 13.22%, 8.57%, 18.71%, and 34.62%, respectively, compared to the untreated sample. The scanning electron microscope observation showed that the pore structure of vegetation concrete was optimized after foam incorporation. Adding plant protein foaming agents to modify the pore structure of vegetation concrete is appropriate, with an optimal foam volume ratio of 20–30%. This study provides new insights and references for slope ecological restoration engineering. Full article
(This article belongs to the Special Issue Functional Cement-Based Composites for Civil Engineering (Volume II))
Show Figures

Figure 1

Figure 1
<p>Vegetation concrete: (<b>a</b>) Vegetation concrete, (<b>b</b>) Consolidation of Vegetation concrete (Magnified view).</p>
Full article ">Figure 2
<p>Foam made by high-speed mixer: (<b>a</b>) High-speed mixer, (<b>b</b>) Foam.</p>
Full article ">Figure 3
<p>Sample preparation: (<b>a</b>) Add powdered components and stir; (<b>b</b>) Add water and stir; (<b>c</b>) Add foam. (<b>d</b>) Stirring, (<b>e</b>) Add larger granular components and stir; (<b>f</b>) Foam vegetation concrete.</p>
Full article ">Figure 4
<p>Vegetation concrete samples with different foam volumes: (<b>a</b>) 0%, (<b>b</b>) 10%, (<b>c</b>) 20%, (<b>d</b>) 30%, (<b>e</b>) 40%, (<b>f</b>) 50%.</p>
Full article ">Figure 5
<p>Change in volume weight and porosity of vegetation concrete with added foam volume.</p>
Full article ">Figure 6
<p>Shear strength of vegetation concrete at different ages.</p>
Full article ">Figure 7
<p>Influence of foam volume on cohesion and angle of internal friction of vegetated concrete.</p>
Full article ">Figure 8
<p>The growth conditions of <span class="html-italic">Festuca arundinacea</span> on vegetation concrete with different treatments: (<b>a</b>) 0%, (<b>b</b>) 10%, (<b>c</b>) 20%, (<b>d</b>) 30%, (<b>e</b>) 40%, (<b>f</b>) 50%.</p>
Full article ">Figure 9
<p>Influence of foam volume on the biomass of <span class="html-italic">Festuca arundinacea</span>.</p>
Full article ">Figure 10
<p>SEM images of vegetation concrete samples at ×2000 magnification: (<b>a</b>) 0%, (<b>b</b>) 10%, (<b>c</b>) 20%, (<b>d</b>) 30%, (<b>e</b>) 40%, (<b>f</b>) 50%.</p>
Full article ">Figure 11
<p>SEM images of vegetation concrete with 0% and 50% foam volume at ×5000 magnification: (<b>a</b>) 0%, (<b>b</b>) 50%.</p>
Full article ">
21 pages, 13380 KiB  
Article
Macro-Mesoscopic Failure Mechanism Based on a Direct Shear Test of a Cemented Sand and Gravel Layer
by Long Qian, Xingwen Guo, Qinghui Liu, Xin Cai and Xiaochuan Zhang
Buildings 2024, 14(12), 4078; https://doi.org/10.3390/buildings14124078 - 23 Dec 2024
Viewed by 434
Abstract
In order to explore the influence of different layer treatment methods on the macro- and meso-mechanical properties of cemented sand and gravel (CSG), in this paper, the shear behavior of CSG material was simulated by a three-dimensional particle flow program (PFC3D) based on [...] Read more.
In order to explore the influence of different layer treatment methods on the macro- and meso-mechanical properties of cemented sand and gravel (CSG), in this paper, the shear behavior of CSG material was simulated by a three-dimensional particle flow program (PFC3D) based on the results of direct shear test in the laboratory. In shear tests, untreated CSG samples with interface coating mortar and chiseling were used, and granular discrete element software (PDC3D 7.0) was used to establish mesoscopic numerical models of CSG samples with the above three interface treatment methods, in order to reveal the effects of interface treatment methods on the interface strength and damage mechanism of CSG samples. The results show that, with the increase in normal stress, the amount of aggregate falling off the shear failure surface increases, the bump and undulation are more obvious, and the failure mode of the test block is inferred to be extrusion friction failure. The shear strength of the mortar interface is 40% higher than that of the untreated interface, and the failure surface is smooth and flat under different normal stresses. The shear strength of the chiseled interface is 10% higher than that of the untreated interface, and the failure surface fluctuates significantly under different normal stresses. Through the analysis of the fracture evolution process in the numerical simulation, it is found that the fracture of the sample at the mortar interface mainly expands along the mortar–aggregate interface and the damage mode is shear slip. However, the cracks of the samples at the gouged interface are concentrated on the upper and lower sides of the interface, and the damage mode is tension–shear. The failure mode of the samples without surface treatment is mainly tensile and shear failure, and the failure mode gradually changes to extrusion friction failure. Full article
Show Figures

Figure 1

Figure 1
<p>Specimen material.</p>
Full article ">Figure 2
<p>Experimental procedure flowchart.</p>
Full article ">Figure 3
<p>Fabrication process of CSG specimens with no treatment forms.</p>
Full article ">Figure 4
<p>Fabrication process of CSG specimens with different interface treatment forms.</p>
Full article ">Figure 5
<p>Direct shear test device.</p>
Full article ">Figure 6
<p>Shear stress–shear displacement curve.</p>
Full article ">Figure 7
<p>Results of interface strength fitting.</p>
Full article ">Figure 8
<p>Interface with no treatment.</p>
Full article ">Figure 9
<p>Damage pattern of spreading mortar specimens.</p>
Full article ">Figure 10
<p>Damage pattern of chiseling specimens.</p>
Full article ">Figure 11
<p>Numerical model of the specimen.</p>
Full article ">Figure 12
<p>Comparison of numerical simulation results with experimental results for different interfaces of treatment.</p>
Full article ">Figure 13
<p>Number of fracture development in the specimen–shear displacement curve.</p>
Full article ">Figure 13 Cont.
<p>Number of fracture development in the specimen–shear displacement curve.</p>
Full article ">Figure 14
<p>Internal fracture distribution during shearing of the specimen.</p>
Full article ">Figure 14 Cont.
<p>Internal fracture distribution during shearing of the specimen.</p>
Full article ">Figure 15
<p>Thickness of fracture distribution at the end of specimen shear test.</p>
Full article ">
16 pages, 3670 KiB  
Article
Impact of Temperature and Humidity on Key Mechanical Properties of Corrugated Board
by Damian Mrówczyński, Tomasz Gajewski, Aram Cornaggia and Tomasz Garbowski
Appl. Sci. 2024, 14(24), 12012; https://doi.org/10.3390/app142412012 - 22 Dec 2024
Viewed by 418
Abstract
This research explores how temperature and relative humidity impact the mechanical properties of corrugated cardboard. Samples were treated under a range of controlled climate conditions in a climate chamber to simulate varying environmental exposures. Following this conditioning, we performed a series of mechanical [...] Read more.
This research explores how temperature and relative humidity impact the mechanical properties of corrugated cardboard. Samples were treated under a range of controlled climate conditions in a climate chamber to simulate varying environmental exposures. Following this conditioning, we performed a series of mechanical tests: the Edge Crush Test (ECT) to assess compressive strength, four-point Bending Tests (BNTs) in both the Machine (MD) and Cross Directions (CD) to evaluate bending stiffness, Sample Torsion Tests (SSTs) for shear stiffness, and Transverse Shear Tests (TSTs) to measure torsional rigidity. By comparing results across these tests, we aim to determine which mechanical property shows the highest sensitivity to changes in humidity levels. Findings from this study are expected to offer valuable insights into the environmental adaptability of corrugated board, particularly for applications in packaging and storage, where climate variability can affect material performance and durability. Such insights will support the development of more robust and adaptable packaging solutions optimised for specific climate conditions. Full article
Show Figures

Figure 1

Figure 1
<p>The measurement sockets with cardboard samples: (<b>a</b>) the bending stiffness test, (<b>b</b>) the edge crush test, (<b>c</b>) the shear stiffness test and (<b>d</b>) the torsion stiffness test.</p>
Full article ">Figure 2
<p>Mechanical tests on cardboards conducted in this study: (<b>a</b>) the bending stiffness test, (<b>b</b>) the edge crush test, (<b>c</b>) the shear stiffness test and (<b>d</b>) the torsion stiffness test.</p>
Full article ">Figure 3
<p>Single set of samples of corrugated boards for all mechanical tests in both the Machine Direction (MD) and the Cross-machine Direction (CD): the Bending stiffness Test (BNT), the Edge Crush Test (ECT), the Shear Stiffness Test (SST), the Torsion Stiffness Test (TST) and thickness.</p>
Full article ">Figure 4
<p>Exemplary experimental results in the form of raw curves illustrating five samples (marked in different colors) data for B flute cardboard at a temperature of 20 °C and a relative humidity of 40% for: (<b>a</b>) the edge crush test, (<b>b</b>) the shear stiffness test, (<b>c</b>) the torsion stiffness test in the cross-machine direction, (<b>d</b>) the torsion stiffness test in the machine direction, (<b>e</b>) the bending stiffness test in the cross-machine direction and (<b>f</b>) the bending stiffness test in the machine direction.</p>
Full article ">
17 pages, 6416 KiB  
Article
Comparative Study of Transverse Shear Characteristics of Shear-Yielding Bolts and Traditional Bolts Based on Numerical Simulations and Direct Shear Tests
by Jianqiang Xu, Xiaohua Yang, Xueming Jia, Haoyu Zhang and Tiangong Zhang
Buildings 2024, 14(12), 4066; https://doi.org/10.3390/buildings14124066 - 21 Dec 2024
Viewed by 523
Abstract
The shear-yielding bolt is a new type of anchoring structure, and its working mechanism in layered rocks is not yet well understood. To investigate its transverse shear characteristics, this paper takes the shear-yielding bolt as the research subject and uses different anchoring states [...] Read more.
The shear-yielding bolt is a new type of anchoring structure, and its working mechanism in layered rocks is not yet well understood. To investigate its transverse shear characteristics, this paper takes the shear-yielding bolt as the research subject and uses different anchoring states of bolts as variables. A comparative study of shear-yielding bolts and traditional bolts is conducted using the Abaqus numerical simulation software and large-scale direct shear tests. The results show that (1) low-modulus material allows a slight displacement between the structural surface layers, which exerts the friction strength between rock mass layers and avoids stress concentration on the bolt. The shear-yielding bolts reach their peak shear stress in the case of greater displacement, averagely increased by 40% compared to traditional anchor bolts. (2) An increase in the moisture content has less influence on the shear-yielding bolt owing to the material properties. When the moisture content of the structural surface rises from 12% to 20%, for cases where the shear-yielding bolts are used, the peak shear stress decreases by 0.12 kPa, which only accounts for 12% of the original strength. (3) There is an optimum thickness of the low-modulus material in the shear-yielding bolt, considering its effect of releasing shear and the bonding effect between it and the bolt. According to the test results and numerical analysis, the optimum thickness is 15 mm. The results of this research provide a reference and basis for future study and engineering applications of shear-yielding bolts. Full article
(This article belongs to the Special Issue Foundation Treatment and Building Structural Performance Enhancement)
Show Figures

Figure 1

Figure 1
<p>Schematic presentation of the structure of a traditional bolt.</p>
Full article ">Figure 2
<p>Schematic presentation of the structure of a shear-yielding bolt.</p>
Full article ">Figure 3
<p>Schematic presentation of experimental specimen.</p>
Full article ">Figure 4
<p>The large-scale direct shear test instrument.</p>
Full article ">Figure 5
<p>Illustration of experimental procedure: (<b>a</b>) specimen assembly; (<b>b</b>) preparation of structural surface material; (<b>c</b>) insertion of low-modulus material; and (<b>d</b>) completion of test assembly.</p>
Full article ">Figure 6
<p>Calculation model and meshing in the simulation: (<b>a</b>) calculation model and (<b>b</b>) meshing.</p>
Full article ">Figure 7
<p>Shear force–displacement curves for normal pressures when moisture content is 16% and normal stress is (<b>a</b>) 0.25 MPa, (<b>b</b>) 0.5 MPa, (<b>c</b>) 0.75 MPa, and (<b>d</b>) 1 MPa.</p>
Full article ">Figure 8
<p>Shear force–displacement curves for shear-yielding bolt anchoring under different moisture contents and low-modulus material thickness conditions. (<b>a</b>) No anchoring; (<b>b</b>) shear-yielding bolt anchoring; and (<b>c</b>) shear-yielding bolt anchoring with different low-modulus material thicknesses.</p>
Full article ">Figure 9
<p>Shear stress distribution of the joint (no anchoring case, normal stress of 0.5 MPa, moisture content of 20%).</p>
Full article ">Figure 10
<p>Stress and plastic deformation distribution of the joint (traditional bolt anchoring, normal stress of 0.5 MPa, moisture content of 16%).</p>
Full article ">Figure 11
<p>Stress and plastic deformation distribution of the joint (shear-yielding bolt, normal stress of 0.5 MPa, moisture content of 16%).</p>
Full article ">Figure 11 Cont.
<p>Stress and plastic deformation distribution of the joint (shear-yielding bolt, normal stress of 0.5 MPa, moisture content of 16%).</p>
Full article ">
16 pages, 11595 KiB  
Article
Experimental and Numerical Simulation Study on the Shear Behavior of Rock-like Specimens with Non-Persistent Joints
by Gang Wang, Hongqi Li and Zhaoying Li
Appl. Sci. 2024, 14(24), 11933; https://doi.org/10.3390/app142411933 - 20 Dec 2024
Viewed by 362
Abstract
Shear failure of non-persistent joints represents a significant contributing factor to rock mass instability. Since non-persistent joints have various parameter characteristics, it is of great practical importance to explore shear behavior with different parameters for preventing geological disasters and engineering construction. In this [...] Read more.
Shear failure of non-persistent joints represents a significant contributing factor to rock mass instability. Since non-persistent joints have various parameter characteristics, it is of great practical importance to explore shear behavior with different parameters for preventing geological disasters and engineering construction. In this study, the effects of joint aperture, joint persistency, and normal stress on the shear behavior of non-persistent persistent joints were investigated by combining indoor tests with numerical simulations. Firstly, an indoor direct shear test was carried out to examine the shear stress, normal displacement, and failure patterns from a macroscopic perspective. Then, a numerical model was constructed using the FEM-CZM method to analyze the stress evolution process of non-persistent joint shear failure from a microscopic perspective. The results show that within the scope of the research, the peak shear strength of non-persistent joints is negatively correlated with joint aperture and joint persistency and positively correlated with normal stress. The residual shear strength is negatively correlated with joint persistency and positively correlated with normal stress. Peak normal displacement is negatively correlated with joint aperture and normal stress, and final normal displacement is negatively correlated with joint persistency and normal stress. The failure pattern of non-persistent joints is affected by internal stress. As joint aperture, joint persistency, and normal stress increase, stress concentration at the rock bridge intensifies, the width of the shear failure zone diminishes, and the specimen changes from tensile failure or mixed failure to shear failure. The research results may enrich the understanding of the shear behavior of non-persistent joints and provide some reference value for safe construction and geological hazard protection. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the direct shear test process and equipment for non-persistent joints: (<b>a</b>) specimens preparation process and non-persistent joint specimens; (<b>b</b>) test instrument; (<b>c</b>) shear box.</p>
Full article ">Figure 2
<p>Shear stress–shear displacement curve: (<b>a</b>) joint aperture; (<b>b</b>) joint persistency; (<b>c</b>) normal stresses.</p>
Full article ">Figure 3
<p>Histograms of peak shear strength and residual shear strength: (<b>a</b>) joint aperture; (<b>b</b>) joint persistency; (<b>c</b>) normal stress.</p>
Full article ">Figure 4
<p>Normal displacement–shear displacement curve: (<b>a</b>) joint aperture; (<b>b</b>) joint persistency; (<b>c</b>) normal stress.</p>
Full article ">Figure 5
<p>Histograms of peak normal displacement and final normal displacement: (<b>a</b>) joint aperture; (<b>b</b>) joint persistency; (<b>c</b>) normal stress.</p>
Full article ">Figure 6
<p>Failure diagram for non-persistent joint specimens: (<b>a</b>) joint aperture; (<b>b</b>) joint persistency; (<b>c</b>) normal stress.</p>
Full article ">Figure 6 Cont.
<p>Failure diagram for non-persistent joint specimens: (<b>a</b>) joint aperture; (<b>b</b>) joint persistency; (<b>c</b>) normal stress.</p>
Full article ">Figure 7
<p>The insertion process of the zero-thickness cohesive elements: (<b>a</b>) adjacent solid elements; (<b>b</b>) discretize solid elements; (<b>c</b>) construct a zero-thickness cohesive element; (<b>d</b>) insert zero-thickness cohesive elements into solid elements.</p>
Full article ">Figure 8
<p>Uniaxial compression test and uniaxial compression numerical model construction: (<b>a</b>) uniaxial compression test; (<b>b</b>) uniaxial compression numerical model construction.</p>
Full article ">Figure 9
<p>Comparison of uniaxial compression test numerical and experimental results.</p>
Full article ">Figure 10
<p>Non-persistent joint finite element numerical model.</p>
Full article ">Figure 11
<p>Comparison of the results of numerical simulations and direct shear tests for non-persistent joints. (<b>a</b>) 30-0.5-2; (<b>b</b>) 30-1.0-2; (<b>c</b>) 30-2.0-2; (<b>d</b>) 20-1.0-2; (<b>e</b>) 40-1.0-2; (<b>f</b>) 30-1.0-1; (<b>g</b>) 30-1.0-3.</p>
Full article ">Figure 11 Cont.
<p>Comparison of the results of numerical simulations and direct shear tests for non-persistent joints. (<b>a</b>) 30-0.5-2; (<b>b</b>) 30-1.0-2; (<b>c</b>) 30-2.0-2; (<b>d</b>) 20-1.0-2; (<b>e</b>) 40-1.0-2; (<b>f</b>) 30-1.0-1; (<b>g</b>) 30-1.0-3.</p>
Full article ">Figure 12
<p>Evolution process of shear stress of non-persistent joints: (<b>a</b>) joint aperture; (<b>b</b>) joint persistency; (<b>c</b>) normal stress.</p>
Full article ">Figure 12 Cont.
<p>Evolution process of shear stress of non-persistent joints: (<b>a</b>) joint aperture; (<b>b</b>) joint persistency; (<b>c</b>) normal stress.</p>
Full article ">
15 pages, 6524 KiB  
Article
A Study on the Effect of Graphene Oxide on Geotechnical Properties of Soil
by Kyungwon Park, Ju-Hoon Kim, Junwoo Shin, Hoyoung Lee and Boo Hyun Nam
Materials 2024, 17(24), 6199; https://doi.org/10.3390/ma17246199 - 18 Dec 2024
Viewed by 294
Abstract
Edge-oxidized graphene oxide (EOGO) is a nano-sized material that is chemically stable and easily mixed with water due to its hydrophilic properties; thus, it has been used in various engineering fields, particularly for the reinforcement of building and construction materials. In this study, [...] Read more.
Edge-oxidized graphene oxide (EOGO) is a nano-sized material that is chemically stable and easily mixed with water due to its hydrophilic properties; thus, it has been used in various engineering fields, particularly for the reinforcement of building and construction materials. In this study, the effect of EOGO in soil reinforcement was investigated. When mixed with soil, it affects the mechanical properties of the soil–GO mixture. Various amounts of the GO (0%, 0.02%, 0.06%, 0.1%) were added into the sand–clay mixture, and their geotechnical properties were evaluated via multiple laboratories testing methods, including a standard Proctor test, direct shear test, compressibility test, and contact angle measurement. The experimental results show that with the addition of EOGO in soil of up to 0.06% EOGO, the compressibility decreases, the shear strength increases, and the maximum dry density (after compaction) increases. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the EOGO manufacturing process [<a href="#B26-materials-17-06199" class="html-bibr">26</a>]; (<b>a</b>) ball-milling and oxidant (<b>b</b>) mechanochemical processes of EOGO production.</p>
Full article ">Figure 2
<p>Particle distribution of the soil mixture (sand and clay).</p>
Full article ">Figure 3
<p>Material characterization of the EOGO: SEM (<b>a</b>) and FTIR (<b>b</b>).</p>
Full article ">Figure 4
<p>Experimental design.</p>
Full article ">Figure 5
<p>Schematic diagram of the direct shear test.</p>
Full article ">Figure 6
<p>Classification of the contact angle measurement between EOGO and the water droplet.</p>
Full article ">Figure 7
<p>Results of the direct shear strength test.</p>
Full article ">Figure 7 Cont.
<p>Results of the direct shear strength test.</p>
Full article ">Figure 8
<p>Comparison of the shear strength of the soil–EOGO mixture determined at various strain levels.</p>
Full article ">Figure 8 Cont.
<p>Comparison of the shear strength of the soil–EOGO mixture determined at various strain levels.</p>
Full article ">Figure 9
<p>Shear strength of the soil–EOGO mixtures with varied EOGO content.</p>
Full article ">Figure 10
<p>Result of the compressibility test (plot of displacement vs. time).</p>
Full article ">Figure 11
<p>Results of the compressibility test: (<b>a</b>) void ratio vs. normal stress; (<b>b</b>) percentage of displacement increment vs. normal stress.</p>
Full article ">Figure 12
<p>Result of the compaction test.</p>
Full article ">Figure 13
<p>Result of the contact angle measurement.</p>
Full article ">
18 pages, 8966 KiB  
Article
Size Effect on Shear Behaviors of 3D-Printed Soft Rock-like Materials Under Different Shear Velocities
by Chaolei Wu, Lishuai Jiang, Yang Zhao, Qi Wu, Yiming Yang, Xiaohan Peng and Pimao Li
Materials 2024, 17(24), 6180; https://doi.org/10.3390/ma17246180 - 18 Dec 2024
Viewed by 338
Abstract
The shear failure of rock masses is one of the primary causes of underground engineering instability. The shear mechanical behavior of rocks at different sizes is of great significance for studying the shear failure pattern of engineering rock masses. However, due to the [...] Read more.
The shear failure of rock masses is one of the primary causes of underground engineering instability. The shear mechanical behavior of rocks at different sizes is of great significance for studying the shear failure pattern of engineering rock masses. However, due to the presence of various joints and defects in natural rocks, the obtained rock specimens exhibit significant discreteness, making it difficult to customize specimen sizes for size effect studies. In recent years, 3D printing (3DP) technology has gained widespread application in rock mechanics tests due to its high printing precision and ability to form specimens in a single step with minimal discreteness. Among these, specimens prepared using sand-powder 3DP exhibit elastoplastic mechanical characteristics similar to those of natural rocks. Therefore, this study utilized sand-powder 3DP to prepare rock-like specimens of four different sizes and conducted compression shear tests under three different shear velocities. The shear strength, shear strain, and wear of the shear surfaces were analyzed as functions of specimen size and shear velocities. The results indicate that under the same shear velocity, the shear strength of the specimens is negatively correlated with specimen size. The peak shear strain is generally unaffected by shear velocities, but it increases initially and then decreases with increasing specimen size. As specimen size increases, the degree of specimen damage intensifies, and larger specimens are more prone to developing derived fractures. This study broadens the application of sand-powder 3DP technology in investigating the shear mechanical properties of soft rocks, offering novel insights into the study of size effects in rock mechanics. However, the current research does not encompass tests on 3D-printed rock specimens with varying printing directions, nor does it delve into the role of fractures in size effect analyses. Future investigations will aim to address these limitations, thereby advancing the applicability of 3D printing technology in rock mechanics research and enhancing its contributions to the field. Full article
(This article belongs to the Special Issue 3D Printing Techniques in Construction Materials)
Show Figures

Figure 1

Figure 1
<p>Flow chart of rock-like specimen prepared by PLA mold [<a href="#B23-materials-17-06180" class="html-bibr">23</a>].</p>
Full article ">Figure 2
<p>A 3D printer and specimen preparation flow chart: (<b>a</b>) Easy3DP-S450 3D printer; (<b>b</b>) BJT process for specimen preparation.</p>
Full article ">Figure 3
<p>MTS816 rock mechanics shear test system and test process.</p>
Full article ">Figure 4
<p>Shear stress–shear displacement curves of typical specimens.</p>
Full article ">Figure 5
<p>Curves of shear stress–shear displacement for sand-powder 3DP specimens at different shear velocities: (<b>a</b>) <span class="html-italic">ν</span> = 0.4 mm/min; (<b>b</b>) <span class="html-italic">ν</span> = 0.8 mm/min; (<b>c</b>) <span class="html-italic">ν</span> = 1.2 mm/min.</p>
Full article ">Figure 6
<p>Bar chart of the relationship between percentage reduction in strength and incremental size increases.</p>
Full article ">Figure 7
<p>Bar chart of shear strength for sand-powder 3DP specimens at different shear velocities.</p>
Full article ">Figure 8
<p>Curves of peak shear strain–shear velocity for sand-powder 3DP soft rock-like specimens of different sizes.</p>
Full article ">Figure 9
<p>Curves of shear modulus and specimen size for sand-powder 3DP soft rock-like specimens of different sizes.</p>
Full article ">Figure 10
<p>Shear failure patterns of sand-powder 3DP soft rock-like specimens with different sizes: (<b>a</b>) specimen size = 40 mm; (<b>b</b>) specimen size = 50 mm; (<b>c</b>) specimen size = 70 mm; (<b>d</b>) specimen size = 100 mm.</p>
Full article ">Figure 11
<p>Three shear failure patterns of sand-powder 3DP soft rock-like specimens: (<b>a</b>) planar failure; (<b>b</b>) step-like failure; (<b>c</b>) cross-like failure.</p>
Full article ">Figure 12
<p>Diagram of the shear failure surface of the specimen.</p>
Full article ">Figure 13
<p>Shear wear conditions and binarized images of specimens of different sizes.</p>
Full article ">Figure 14
<p>Shear stress–shear displacement curves of sand powder 3DP rock-like and natural rocks.</p>
Full article ">
19 pages, 9294 KiB  
Article
Study on Interlayer Interface Deterioration of Double-Block Ballastless Track in Humid and Hot Environments Based on Acoustic Emission Technique
by Yuchen Luo, Yuhang Liu and Siming Liang
Buildings 2024, 14(12), 3997; https://doi.org/10.3390/buildings14123997 - 17 Dec 2024
Viewed by 440
Abstract
The deterioration of the interlayer interface of a double-block ballastless track is affected by the environmental temperature and moisture conditions, which will have a negative effect on its service life. Composite specimens with interlayer interfaces of double-block ballastless track were fabricated and deteriorated [...] Read more.
The deterioration of the interlayer interface of a double-block ballastless track is affected by the environmental temperature and moisture conditions, which will have a negative effect on its service life. Composite specimens with interlayer interfaces of double-block ballastless track were fabricated and deteriorated by an accelerated method, i.e., immersed in saturated ammonium chloride solution with various temperatures for different times. Then, the deterioration condition and mechanical properties of the composite specimens were investigated experimentally by a universal material testing machine and acoustic emission technique. The automatic sensor test (AST) method is capable of assessing the deterioration condition of the interlayer interface based on the relative wave velocity. The deterioration depth of the interlayer interface tends to increase with increasing solution temperature and immersion time. Both the solution temperature and immersion time have a negative impact on the splitting tensile strength and direct shear strength. A linear relation is found between the splitting tensile strength (direct shear strength) and the cumulative AE energy released at the fracture moment. The damage factor defined by the cumulative AE energy for most composite specimens is no greater than 0.2 before they are going to be fractured but increases sharply to 1.0 at the fracture moment. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the double-block ballastless track.</p>
Full article ">Figure 2
<p>Illustration of the accelerated deterioration test under humid and hot conditions.</p>
Full article ">Figure 3
<p>(<b>a</b>) Measurement of the wave velocity of the composite specimen by the AST method, and (<b>b</b>) illustration of the positions of the AE sensors.</p>
Full article ">Figure 4
<p>Illustration of (<b>a</b>) splitting tensile test and (<b>b</b>) direct shear test of composite specimens.</p>
Full article ">Figure 5
<p>Devices used in the interface bonding performance tests of composite specimens.</p>
Full article ">Figure 6
<p>Measured relative wave velocities at different positions of the composite specimens.</p>
Full article ">Figure 7
<p>Load–displacement curves of composite specimens during the splitting tensile test.</p>
Full article ">Figure 8
<p>Measured splitting tensile strength under various humid and hot conditions.</p>
Full article ">Figure 9
<p>Load–displacement curves of composite specimens during the direct shear test.</p>
Full article ">Figure 10
<p>Measured direct shear strength under various humid and hot conditions.</p>
Full article ">Figure 11
<p>Measured AE energy evolution of composite specimen after 7-day deterioration during the splitting tensile test.</p>
Full article ">Figure 11 Cont.
<p>Measured AE energy evolution of composite specimen after 7-day deterioration during the splitting tensile test.</p>
Full article ">Figure 12
<p>Measured AE energy evolution of composite specimen after 14-day deterioration during the splitting tensile test.</p>
Full article ">Figure 13
<p>Relationship between splitting tensile strength and cumulative AE energy.</p>
Full article ">Figure 14
<p>Measured AE energy evolution of composite specimen after 7-day deterioration during the direct shear test.</p>
Full article ">Figure 15
<p>Measured AE energy evolution of composite specimen after 14-day deterioration during the direct shear test.</p>
Full article ">Figure 15 Cont.
<p>Measured AE energy evolution of composite specimen after 14-day deterioration during the direct shear test.</p>
Full article ">Figure 16
<p>Relationship between direct shear strength and cumulative AE energy.</p>
Full article ">Figure 17
<p>Relationship between damage factor and normalized splitting tensile load of composite specimens after (<b>a</b>) 7-day and (<b>b</b>) 14-day deterioration.</p>
Full article ">Figure 18
<p>Relationship between damage factor and normalized shear load of composite specimens after (<b>a</b>) 7-day and (<b>b</b>) 14-day deterioration.</p>
Full article ">
Back to TopTop