[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (17)

Search Parameters:
Keywords = cohesive admixture

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 12544 KiB  
Article
Evaluation of Shear Strength and Stiffness of a Loess–Sand Mixture in Triaxial and Unconfined Compression Tests
by Matylda Tankiewicz, Magdalena Kowalska and Jakub Mońka
Materials 2024, 17(15), 3831; https://doi.org/10.3390/ma17153831 - 2 Aug 2024
Viewed by 390
Abstract
Mechanical soil parameters are not constants and can be defined in various ways. Therefore, determination of their values for engineering practice is difficult. This problem is discussed based on results of piezoceramic element tests and triaxial tests (unconfined and confined) on loess specimens [...] Read more.
Mechanical soil parameters are not constants and can be defined in various ways. Therefore, determination of their values for engineering practice is difficult. This problem is discussed based on results of piezoceramic element tests and triaxial tests (unconfined and confined) on loess specimens improved by compaction and sand admixture (20% by weight). The study indicated also the effectiveness of this simple method of loess stabilization. The influence of specimen size, draining conditions, stress and strain state, and different calculation methods on the evaluation of basic mechanical parameters were analyzed. The initial shear and Young’s moduli, the degradation of secant moduli with strain, tangent moduli, and Poisson’ ratio were determined. The results showed that the shear strength parameters are much less sensitive to the test variables than the stiffness parameters are. In triaxial tests, the strength criterion adopted, the sample size, and the drainage conditions influenced the measured value of cohesion, with a much smaller impact on the angle of internal friction. On the other hand, the adopted definition of the parameter and the range of strains had the greatest influence on the value of the stiffness modulus. Moreover, larger specimens were usually found to be stiffer. Full article
(This article belongs to the Special Issue Experimental Tests and Numerical Analysis of Construction Materials)
Show Figures

Figure 1

Figure 1
<p>Microscopic image of the tested loess (ZEISS SteREO Discovery.V20 stereomicroscope, ZEISS, Oberkochen, Germany).</p>
Full article ">Figure 2
<p>Grain size distribution of loess, sand, and loess–sand mixture.</p>
Full article ">Figure 3
<p>Shear characteristics for six UC/38 (<b>a</b>) and six UC/70 (<b>b</b>) specimens based on external sensors (black dashed lines) and local sensors (red continuous lines).</p>
Full article ">Figure 4
<p>Poisson’s ratio values for (<b>a</b>) UC/38 and (<b>b</b>) UC/70 specimens based on local sensors.</p>
Full article ">Figure 5
<p>Stress paths (<b>a</b>) and shear characteristics (<b>b</b>) for CIU/38 and CIU/70 specimens.</p>
Full article ">Figure 6
<p>Stress paths (<b>a</b>) and shear characteristics (<b>b</b>) for CID/38 and CID/70 specimens.</p>
Full article ">Figure 7
<p>Influence of the mean effective stress <span class="html-italic">p’</span><sub>0</sub> on (<b>a</b>) the initial shear modulus <span class="html-italic">G</span><sub>0</sub> and (<b>b</b>) the initial Young’s modulus <span class="html-italic">E</span><sub>0</sub> [<a href="#B9-materials-17-03831" class="html-bibr">9</a>].</p>
Full article ">Figure 8
<p>Evolution of the Poisson’s ratio <span class="html-italic">ν</span><sub>0</sub> (Equation (9)) with the mean effective stress <span class="html-italic">p’</span><sub>0</sub>.</p>
Full article ">Figure 9
<p>Degradation curves of the normalized secant shear modulus <span class="html-italic">G<sub>s</sub></span>/<span class="html-italic">G</span><sub>0</sub> from (<b>a</b>) undrained CIU/38 and (<b>b</b>) drained CID/38 triaxial tests.</p>
Full article ">Figure 10
<p>Degradation curves of the normalized tangent shear modulus <span class="html-italic">G<sub>t</sub></span>/<span class="html-italic">G</span><sub>0</sub> from (<b>a</b>) undrained CIU/38 and (<b>b</b>) drained CID/38 triaxial tests.</p>
Full article ">Figure 11
<p>Degradation curves of the normalized secant <span class="html-italic">G<sub>s</sub></span>/<span class="html-italic">G</span><sub>0</sub> shear modulus from drained CID/38 (<b>a</b>) and undrained CIU/38 (<b>b</b>) tests.</p>
Full article ">Figure 12
<p>Comparison of normalized secant <span class="html-italic">G<sub>s</sub></span>/<span class="html-italic">G</span><sub>0</sub> and tangent <span class="html-italic">G<sub>t</sub></span>/<span class="html-italic">G</span><sub>0</sub> shear modulus from undrained CIU/38 (<b>a</b>) and undrained CID/38 (<b>b</b>) tests.</p>
Full article ">Figure 13
<p>Comparison of the degradation curves <span class="html-italic">G<sub>s</sub></span>/<span class="html-italic">G</span><sub>0</sub> for specimens with different sizes (38 and 70 mm in diameter) in undrained CIU (<b>a</b>) and drained CID (<b>b</b>) triaxial tests.</p>
Full article ">Figure 14
<p>Comparison of the degradation curves <span class="html-italic">G<sub>s</sub></span>/<span class="html-italic">G</span><sub>0</sub> (<b>a</b>) and <span class="html-italic">G<sub>t</sub></span>/<span class="html-italic">G</span><sub>0</sub> (<b>b</b>) for specimens sheared at <span class="html-italic">σ’</span><sub>3</sub> = const (conventional stress path) and at <span class="html-italic">p’</span> = const (vertical stress path).</p>
Full article ">Figure 15
<p>Evolution of the Poisson’s ratio <span class="html-italic">ν</span> with stress ratio <span class="html-italic">SF</span> = <span class="html-italic">q</span>/(<span class="html-italic">q)<sub>max</sub></span>—CID/38 specimens.</p>
Full article ">Figure 16
<p>Degradation curves of the normalized secant Young’s modulus <span class="html-italic">E<sub>s</sub></span>/<span class="html-italic">E</span><sub>0</sub> from undrained CIU/38 (<b>a</b>) and drained CID/38 (<b>b</b>) triaxial tests.</p>
Full article ">Figure 17
<p>Degradation curves of the normalized tangent Young’s modulus <span class="html-italic">E<sub>t</sub></span>/<span class="html-italic">E</span><sub>0</sub> from undrained CIU/38 (<b>a</b>) and drained CID/38 (<b>b</b>) triaxial tests.</p>
Full article ">Figure 18
<p>Influence of the consolidation pressure <span class="html-italic">σ’<sub>C</sub></span> on <span class="html-italic">E</span><sub>50</sub> values for specimens 38 mm (<b>a</b>) and 70 mm (<b>b</b>) in diameter.</p>
Full article ">
19 pages, 6919 KiB  
Article
Study of the Performance of Emulsified Asphalt Shotcrete in High-Altitude Permafrost Regions
by Yitong Hou, Kaimin Niu, Bo Tian, Xueyang Li and Junli Chen
Coatings 2024, 14(6), 692; https://doi.org/10.3390/coatings14060692 - 1 Jun 2024
Viewed by 384
Abstract
To improve the performance of shotcrete in high-altitude and low-temperature environments, emulsified asphalt shotcrete (EASC), which can be used in negative-temperature environments, was prepared by using low-freezing-point emulsified asphalt, calcium aluminate cement, and sodium pyrophosphate as modified materials. The effect of emulsified asphalt [...] Read more.
To improve the performance of shotcrete in high-altitude and low-temperature environments, emulsified asphalt shotcrete (EASC), which can be used in negative-temperature environments, was prepared by using low-freezing-point emulsified asphalt, calcium aluminate cement, and sodium pyrophosphate as modified materials. The effect of emulsified asphalt on the performance of shotcrete was investigated through concrete spraying and indoor tests. Then, the modification mechanism of emulsified asphalt with respect to EASC was analyzed by combining scanning electron microscopy images and the pore structure characteristics of EASC. The results showed that in a negative-temperature environment, the incorporation of emulsified asphalt delayed the formation of the peak of the cement hydration exotherm, slowed the rate of the cement hydration exotherm, reduced the thermal perturbation of permafrost by EASC, increased the cohesion of the concrete, improved the bond strength between EASC and permafrost, and reduced the rate of rebound. The mechanical strength of the studied EASC decreased upon increasing the amount of emulsified asphalt in the admixture, and its resistance to cracking gradually improved. A content of less than 5% emulsified asphalt could improve the internal pore structure of EASC, thus improving its durability. Increasing the content of emulsified asphalt affected the hydration process of the cement, and the volume content of the capillary pores and macropores increased, which reduced the durability of the EASC. Full article
(This article belongs to the Special Issue Novel Cleaner Materials for Pavements)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of the cement: (<b>a</b>) Portland cement; (<b>b</b>) calcium aluminate cement.</p>
Full article ">Figure 2
<p>Optical microscope photograph of low-freezing-point emulsified asphalt.</p>
Full article ">Figure 3
<p>Constant-temperature and constant-humidity environment test chamber.</p>
Full article ">Figure 4
<p>Shotcrete hydration temperature rise test. (<b>a</b>) experimental design; and (<b>b</b>) experimental process.</p>
Full article ">Figure 5
<p>Split tensile test. (<b>a</b>) experimental design; and (<b>b</b>) experimental process.</p>
Full article ">Figure 6
<p>Principle of the concrete spraying test.</p>
Full article ">Figure 7
<p>Concrete spraying test process: (<b>a</b>) the spraying process; (<b>b</b>) the large-plate after spraying was carried out.</p>
Full article ">Figure 8
<p>Thermal disturbance of permafrost by EASC: (<b>a</b>) at the interface and (<b>b</b>) at a depth of 3 cm in frozen soil.</p>
Full article ">Figure 9
<p>EASC bond strength.</p>
Full article ">Figure 10
<p>EASC rebound rate test results.</p>
Full article ">Figure 11
<p>EASC mechanical strength: (<b>a</b>) compressive strength; (<b>b</b>) flexural strength; and (<b>c</b>) flexural/compressive ratio.</p>
Full article ">Figure 12
<p>EASC electric flux test results.</p>
Full article ">Figure 13
<p>EASC freeze–thaw cycle test results: (<b>a</b>) mass loss rate and (<b>b</b>) relative dynamic elastic modulus.</p>
Full article ">Figure 14
<p>SEM images and energy spectrum analyses of EASC specimens: (<b>a</b>) EA-0; (<b>b</b>) EA-2.5; (<b>c</b>) EA-5; (<b>d</b>) EA-7.5; and (<b>e</b>) EA-10.</p>
Full article ">Figure 14 Cont.
<p>SEM images and energy spectrum analyses of EASC specimens: (<b>a</b>) EA-0; (<b>b</b>) EA-2.5; (<b>c</b>) EA-5; (<b>d</b>) EA-7.5; and (<b>e</b>) EA-10.</p>
Full article ">Figure 15
<p>MIP test results for EASC: (<b>a</b>) pore size distribution differential curve and (<b>b</b>) pore size distribution.</p>
Full article ">
19 pages, 4449 KiB  
Article
The Application of Sand Transport with Cohesive Admixtures Model for Predicting Flushing Flows in Channels
by Leszek M. Kaczmarek, Jerzy Zawisza, Iwona Radosz, Magdalena Pietrzak and Jarosław Biegowski
Water 2024, 16(9), 1214; https://doi.org/10.3390/w16091214 - 24 Apr 2024
Cited by 1 | Viewed by 640
Abstract
The feature of self-cleansing in sewer pipes is a standard requirement in the design of drainage systems, as sediments deposited on the channel bottom cause changes in channel geometric properties and in hydrodynamic parameters, including the friction caused by the cohesive forces of [...] Read more.
The feature of self-cleansing in sewer pipes is a standard requirement in the design of drainage systems, as sediments deposited on the channel bottom cause changes in channel geometric properties and in hydrodynamic parameters, including the friction caused by the cohesive forces of sediment fractions. Here, it is shown that the content of cohesive fractions significantly inhibits the transport of non-cohesive sediments. This paper presents an advanced calculation procedure for estimating flushing flows in channels. This procedure is based on innovative predictive models developed for non-cohesive and granulometrically heterogeneous sediment transport with additional cohesive fraction content to estimate the magnitude of increased flow necessary to ensure self-cleansing of channels. The computations according to the proposed procedure were carried out for a wide range of hydrodynamic conditions, two grain diameters, six cohesive (clay) fraction additive contents and two critical stress values. The trend lines of calculations were composed with the results of experimental studies in hydraulic flumes. Full article
Show Figures

Figure 1

Figure 1
<p>Vertical structure of sediment transport and the layer <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>z</mi> </mrow> </semantics></math> of erosion.</p>
Full article ">Figure 2
<p>Flow charts of numerical algorithms for calculations of sediment transport with/without cohesion by model [<a href="#B44-water-16-01214" class="html-bibr">44</a>] and [<a href="#B42-water-16-01214" class="html-bibr">42</a>,<a href="#B43-water-16-01214" class="html-bibr">43</a>], respectively; whereby: <math display="inline"><semantics> <mrow> <msubsup> <mi>τ</mi> <mo>∗</mo> <mo>′</mo> </msubsup> </mrow> </semantics></math>—the skin shear stress at the top of the contact layer; <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi>i</mi> </msub> </mrow> </semantics></math>—content of grains with diameter <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mi>i</mi> </msub> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mi>r</mi> </msub> </mrow> </semantics></math>—representative diameter in grain collision sub-layer; <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mn>0</mn> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mrow> <mn>0</mn> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>—the shear stress at the top of grain collision sub-layer; <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mn>0</mn> </msub> </mrow> </semantics></math>—without cohesion; <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mrow> <mn>0</mn> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>—with cohesion; <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mi>g</mi> </msub> </mrow> </semantics></math>—thickness of grain collision sub-layer; <math display="inline"><semantics> <mrow> <msubsup> <mi>u</mi> <mrow> <mi>f</mi> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> </mrow> </semantics></math>—the shear velocity due to cohesion; <math display="inline"><semantics> <mi>N</mi> </semantics></math>—number of fraction; <math display="inline"><semantics> <mi>h</mi> </semantics></math>—water depth; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> </mrow> </semantics></math>—critical Shields parameter; MPM formula—Meyer-Peter and Muller formula 1948 [<a href="#B14-water-16-01214" class="html-bibr">14</a>].</p>
Full article ">Figure 3
<p>Graphical representation of the assumption in the calculation of <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>h</mi> </msub> <mo>=</mo> <mi>h</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mn>0</mn> </msub> <mo>=</mo> <mi>B</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Flow charts of numerical algorithms for calculations of flushing flows <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mi>k</mi> </msub> </mrow> </semantics></math> where: <math display="inline"><semantics> <mi>n</mi> </semantics></math>—is Maning’s roughness <math display="inline"><semantics> <mrow> <mfenced close="]" open="["> <mrow> <msup> <mi>m</mi> <mrow> <mfrac bevelled="true"> <mn>1</mn> <mn>3</mn> </mfrac> </mrow> </msup> <msup> <mi>s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </mfenced> </mrow> </semantics></math>, <math display="inline"><semantics> <mi>I</mi> </semantics></math>—slope of the bottom <math display="inline"><semantics> <mrow> <mfenced close="]" open="["> <mo>−</mo> </mfenced> </mrow> </semantics></math>; Kaczmarek et al. 2019 [<a href="#B41-water-16-01214" class="html-bibr">41</a>]; Zawisza et al. 2023a [<a href="#B42-water-16-01214" class="html-bibr">42</a>]; Zawisza et al. 2023b [<a href="#B43-water-16-01214" class="html-bibr">43</a>].</p>
Full article ">Figure 5
<p>Dependence <math display="inline"><semantics> <mrow> <mi>θ</mi> <msub> <mo>′</mo> <mo>∗</mo> </msub> <mi>c</mi> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <mi>η</mi> <mo> </mo> <mfenced close="]" open="["> <mo>%</mo> </mfenced> </mrow> </semantics></math> for the results of experiments Gdansk 2021 and Ghent 1998 adopted for calculations.</p>
Full article ">Figure 6
<p>Dependence <math display="inline"><semantics> <mrow> <msub> <mo>Φ</mo> <mi>s</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mrow> <msup> <mi>θ</mi> <mo>′</mo> </msup> </mrow> <mo>∗</mo> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>5</mn> <mo>%</mo> </mrow> </semantics></math>—Gdansk 2021 data; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msubsup> <mi>θ</mi> <mrow> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> <mo>=</mo> </mrow> </semantics></math>0.0025.</p>
Full article ">Figure 7
<p>(<b>a</b>) Dependence<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mo>Φ</mo> <mi>s</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mrow> <msup> <mi>θ</mi> <mo>′</mo> </msup> </mrow> <mo>∗</mo> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>10</mn> <mo>%</mo> </mrow> </semantics></math>—Gdansk 2021 data; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msubsup> <mi>θ</mi> <mrow> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> <mo>=</mo> </mrow> </semantics></math> 0.0035. (<b>b</b>) Dependence<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mo>Φ</mo> <mi>s</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mrow> <msup> <mi>θ</mi> <mo>′</mo> </msup> </mrow> <mo>∗</mo> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>10</mn> <mo>%</mo> </mrow> </semantics></math>—Ghent 1998 data; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>0.21</mn> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msubsup> <mi>θ</mi> <mrow> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> <mo>=</mo> <mn>0.004</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7 Cont.
<p>(<b>a</b>) Dependence<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mo>Φ</mo> <mi>s</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mrow> <msup> <mi>θ</mi> <mo>′</mo> </msup> </mrow> <mo>∗</mo> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>10</mn> <mo>%</mo> </mrow> </semantics></math>—Gdansk 2021 data; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msubsup> <mi>θ</mi> <mrow> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> <mo>=</mo> </mrow> </semantics></math> 0.0035. (<b>b</b>) Dependence<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mo>Φ</mo> <mi>s</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mrow> <msup> <mi>θ</mi> <mo>′</mo> </msup> </mrow> <mo>∗</mo> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>10</mn> <mo>%</mo> </mrow> </semantics></math>—Ghent 1998 data; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>0.21</mn> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msubsup> <mi>θ</mi> <mrow> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> <mo>=</mo> <mn>0.004</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Dependence<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mo>Φ</mo> <mi>s</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mrow> <msup> <mi>θ</mi> <mo>′</mo> </msup> </mrow> <mo>∗</mo> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>15</mn> <mo>%</mo> </mrow> </semantics></math>—Gdansk 2021 data; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msubsup> <mi>θ</mi> <mrow> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> <mo>=</mo> <mn>0.0045</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>(<b>a</b>) Dependence<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mo>Φ</mo> <mi>s</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mrow> <msup> <mi>θ</mi> <mo>′</mo> </msup> </mrow> <mo>∗</mo> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>20</mn> <mo>%</mo> </mrow> </semantics></math>—Gdansk 2021 data; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msubsup> <mi>θ</mi> <mrow> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> <mo>=</mo> <mn>0.006</mn> </mrow> </semantics></math>. (<b>b</b>) Dependence<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mo>Φ</mo> <mi>s</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mrow> <msup> <mi>θ</mi> <mo>′</mo> </msup> </mrow> <mo>∗</mo> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>20</mn> <mo>%</mo> </mrow> </semantics></math>—Ghent 1998 data; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>0.21</mn> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msubsup> <mi>θ</mi> <mrow> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> <mo>=</mo> <mn>0.009</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>Dependence<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mo>Φ</mo> <mi>s</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mrow> <msup> <mi>θ</mi> <mo>′</mo> </msup> </mrow> <mo>∗</mo> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>30</mn> <mo>%</mo> </mrow> </semantics></math>—Ghent 1998 data; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>0.21</mn> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msubsup> <mi>θ</mi> <mrow> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> <mo>=</mo> <mn>0.0125</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>Dependence<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mo>Φ</mo> <mi>s</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mrow> <msup> <mi>θ</mi> <mo>′</mo> </msup> </mrow> <mo>∗</mo> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>30</mn> <mo>%</mo> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msubsup> <mi>θ</mi> <mrow> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> <mo>=</mo> <mn>0.0125</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 12
<p>Dependence<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mo>Φ</mo> <mi>s</mi> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mrow> <msup> <mi>θ</mi> <mo>′</mo> </msup> </mrow> <mo>∗</mo> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>30</mn> <mo>%</mo> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>0.21</mn> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msubsup> <mi>θ</mi> <mrow> <mo>∗</mo> <mi>c</mi> </mrow> <mo>′</mo> </msubsup> <mo>=</mo> <mn>0.0125</mn> </mrow> </semantics></math>.</p>
Full article ">
21 pages, 11771 KiB  
Article
Mechanical and Microstructural Changes in Expansive Soils Treated with Lime and Lignin Fiber from Paper Industry
by Taian Wang and Yejiao Wang
Appl. Sci. 2024, 14(8), 3393; https://doi.org/10.3390/app14083393 - 17 Apr 2024
Cited by 1 | Viewed by 664
Abstract
Expansive soil exhibits significant swellings and shrinkages, which may result in severe damage or the collapse of structures built upon it. Calcium-based admixtures, such as lime, are commonly used to improve this problematic soil. However, traditional chemical additions can increase significant environmental stress. [...] Read more.
Expansive soil exhibits significant swellings and shrinkages, which may result in severe damage or the collapse of structures built upon it. Calcium-based admixtures, such as lime, are commonly used to improve this problematic soil. However, traditional chemical additions can increase significant environmental stress. This paper proposes a sustainable solution, namely, the use of lignin fiber (LF) from the paper industry to partially replace lime as an amendment for expansive soils. Both the macroscopic and microscopic characteristics of the lignin fiber-treated expansive soil are extensively studied. The results show that the mechanical properties of expansive soil are improved by using lignin fiber alone. Under the condition of an optimal dosage of 8%, the compressive strength of lignin fiber-modified soil can reach 193 kPa, the shear strength is increased by 40% compared with the untreated soil, and the water conductivity is also improved with the increase in dosage. In addition, compared with 2% lime-modified soil, the compressive strength of 8% lignin fiber- and 2% lime composite-treated expansive soil increased by 50%, the cohesion increased by 12%, and the water conductivity decreased significantly. The microstructure analysis shows that at an 8% lignin fiber content, lignin fibers interweave into a network in the soil, which effectively enhances the strength and stability of the improved soil. Simultaneously, the fibers can form bridges across the adjacent micropores, leading to the merging of pores and transforming fine, dispersed micropores into larger, connected macropores. Lime promotes the flocculation of soil particles, forming larger aggregates and thus resulting in larger pores. The addition of fibers exerts an inhibitory effect on the flocculation reaction in the composite-improved soil. In conclusion, lignin fibers are an effective addition used to partially replace calcium admixture for the treatment of expansive soil, which provides a sustainable and environmentally friendly treatment scheme for reducing industrial waste. Full article
Show Figures

Figure 1

Figure 1
<p>Particle gradation curve of Ankang expansive soil.</p>
Full article ">Figure 2
<p>Tested materials: (<b>a</b>) lignin fiber and (<b>b</b>) lime.</p>
Full article ">Figure 3
<p>The relationship curves between the axial stress and axial strain of the expansive soil improved by different contents of lignin fibers.</p>
Full article ">Figure 4
<p>The relationship curves of the shear stress and horizontal displacement of the expansive soil improved by different contents of lignin fibers.</p>
Full article ">Figure 5
<p>The fitting curves of the shear strength and vertical stress of the expansive soil improved by different contents of lignin fibers.</p>
Full article ">Figure 6
<p>The relationship curve of the hydraulic conductivity and lignin fiber contents of the expansive soil improved by lignin fibers.</p>
Full article ">Figure 7
<p>Cumulative pore distribution curves of expansive soil improved by different contents of lignin fibers.</p>
Full article ">Figure 8
<p>The pore density distribution curves of expansive soil improved by different contents of lignin fibers.</p>
Full article ">Figure 9
<p>The percentage change in various types of pores in expansive soil improved by different contents of lignin fibers.</p>
Full article ">Figure 10
<p>Scanning electron microscopy of lignin fiber-improved soil: (<b>a</b>) lignin fiber content of 0% (500 times), (<b>b</b>) lignin fiber content of 2% (500 times), (<b>c</b>) lignin fiber content of 8% (500 times), and (<b>d</b>) lignin fiber content of 8% (300 times) (different magnifications for subfigure (<b>d</b>)).</p>
Full article ">Figure 11
<p>The relationship between the lignin fiber contents and the compressive strength of the composite-improved soil.</p>
Full article ">Figure 12
<p>The relationship between the compressive strength and the curing days of the improved soils with different treatments.</p>
Full article ">Figure 13
<p>The relationship between the shear stress and the horizontal displacement of the improved soils with different treatments (200 kPa).</p>
Full article ">Figure 14
<p>The fitting curves of the shear stress and vertical stress of the improved soils with different treatments.</p>
Full article ">Figure 15
<p>The relationship between the cohesions and the curing days of the improved soils with different treatments.</p>
Full article ">Figure 16
<p>The relationship between the internal friction angle and the curing days of the improved soils with different treatments.</p>
Full article ">Figure 17
<p>The relationship between the hydraulic conductivity and the curing days of the improved soils with different treatments.</p>
Full article ">Figure 18
<p>Cumulative pore distribution curves of the improved soils with different treatments.</p>
Full article ">Figure 19
<p>The pore density distribution curves of the improved soils with different treatments.</p>
Full article ">Figure 20
<p>Changes in the percentage of various pore types in the expansive soil improved by different treatments.</p>
Full article ">Figure 21
<p>Scanning electron microscopy of composite-improved soils with different contents of lignin fibers and 2% lime: (<b>a</b>) lignin fiber content of 0% (1000 times), (<b>b</b>) lignin fiber content of 2% (1000 times), and (<b>c</b>) lignin fiber content of 8% (1000 times).</p>
Full article ">
17 pages, 9513 KiB  
Article
Analysis of the Effect of Ultra-Fine Cement on the Microscopic Pore Structure of Cement Soil in a Peat Soil Environment
by Jing Cao, Chenhui Huang, Huafeng Sun, Yongfa Guo, Wenyun Ding and Guofeng Hua
Appl. Sci. 2023, 13(23), 12700; https://doi.org/10.3390/app132312700 - 27 Nov 2023
Viewed by 850
Abstract
Treating peat soil foundations around Dianchi Lake and Erhai Lake in Yunnan is a complex problem in practical engineering projects. Peat soil solely reinforced with ordinary cement (OPC) does not satisfy demand. This study aims to solidify soil to achieve better mechanical properties. [...] Read more.
Treating peat soil foundations around Dianchi Lake and Erhai Lake in Yunnan is a complex problem in practical engineering projects. Peat soil solely reinforced with ordinary cement (OPC) does not satisfy demand. This study aims to solidify soil to achieve better mechanical properties. The preparation of peat soil incorporates a humic acid (HA) reagent into cohesive soil, and cement and ultra-fine cement (UFC) are mixed by stirring to prepare cement soil samples. They are then immersed in fulvic acid (FA) solution to simulate cement soil in the actual environment. X-ray diffraction (XRD), mercury intrusion porosimetry (MIP), scanning electron microscopy (SEM), and pores and cracks analysis system (PCAS) tests are used to study the impact of the UFC on the microscopic pore structure of cement soil in a peat soil environment. The unconfined compressive strength (UCS) test is used for verification. The microscopic test results indicate that incorporating UFC enhances the specimen’s micropore structure. The XRD test results show the presence of C–S–H, C–A–S–H, and C–A–H. SEM and PCAS tests show that the UFC proportion increases by between 0% and 10%, and the percentage reduction in the macropore volume is the largest, at 38.84%. When the UFC admixture is 30%, the cumulative reduction in the percentage of macropore volume reaches 71.55%. The MIP test results show that the cumulative volume greater than 10 µm in pore size decreases from 7.68% to 0.17% with an increase in the UFC proportion. The UCS test results show that the maximum strength growth of cement soil is 12.99% when the UFC admixture is 0–10%. Incorporating UFC to form a compound curing agent solves the problem of the traditional reinforcement treatment of peat soil foundation being undesirable and decreases the amount of cement. This study provides practical guidance for reducing carbon emissions in actual projects. Full article
Show Figures

Figure 1

Figure 1
<p>Test material diagram.</p>
Full article ">Figure 2
<p>(<b>a</b>) Microstructure image of the cohesive soil aggregate [<a href="#B30-applsci-13-12700" class="html-bibr">30</a>]. (<b>b</b>) Microstructure images of the HA aggregate [<a href="#B30-applsci-13-12700" class="html-bibr">30</a>].</p>
Full article ">Figure 3
<p>Microstructure image of the FA aggregate.</p>
Full article ">Figure 4
<p>(<b>a</b>) The cumulative distribution curve of the OPC and UFC grain size gradation. (<b>b</b>) OPC and UFC grain size distribution curve.</p>
Full article ">Figure 5
<p>Test procedure.</p>
Full article ">Figure 6
<p>XRD pattern of the test soil (cohesive soil).</p>
Full article ">Figure 7
<p>X-ray diffraction patterns of specimens with various UFC proportions (HA 15%, pH = 6.0, 90 d).</p>
Full article ">Figure 8
<p>Pore diameter distribution curves of cement soil with different UFC proportions. (<b>a</b>) The distribution curve of pore volume percentage. (<b>b</b>) Accumulation curve of pore volume percentage greater than a certain diameter (HA 15%, pH = 6.0, 90 d).</p>
Full article ">Figure 9
<p>Microstructure images of cement soil with various UFC proportions [<a href="#B19-applsci-13-12700" class="html-bibr">19</a>] and PCAS pore segmentation processing images. (<b>a</b>) UFC proportion 0%. (<b>b</b>) UFC proportion 10%. (<b>c</b>) UFC proportion 20%. (<b>d</b>) UFC proportion 30%. (<b>e</b>) UFC proportion 40%. (<b>f</b>) UFC proportion 50%. (1) 500 times SEM image. (2) 2000 times SEM image of yellow frame. (3) PCAS pore segmentation image.</p>
Full article ">Figure 9 Cont.
<p>Microstructure images of cement soil with various UFC proportions [<a href="#B19-applsci-13-12700" class="html-bibr">19</a>] and PCAS pore segmentation processing images. (<b>a</b>) UFC proportion 0%. (<b>b</b>) UFC proportion 10%. (<b>c</b>) UFC proportion 20%. (<b>d</b>) UFC proportion 30%. (<b>e</b>) UFC proportion 40%. (<b>f</b>) UFC proportion 50%. (1) 500 times SEM image. (2) 2000 times SEM image of yellow frame. (3) PCAS pore segmentation image.</p>
Full article ">Figure 10
<p>PCAS test results (percentage of macropore volume).</p>
Full article ">Figure 11
<p>UCS curves of cement soil with various UFC proportions.</p>
Full article ">Figure 12
<p>The strength growth rate of specimens within various UFC proportion ranges.</p>
Full article ">
26 pages, 11932 KiB  
Article
Experimental Study and Finite Element Analysis on the Modification of Fast-Hardening Polymer Cement Composite Material Applied to the Anchorage Zone of Expansion Joint
by Hang Sun, Huan Yuan, Yongming Sun, Xi Li and Liang Luo
Buildings 2023, 13(12), 2910; https://doi.org/10.3390/buildings13122910 - 22 Nov 2023
Viewed by 907
Abstract
Bridges’ expansion joints are prone to damage during operation, and repairing them often requires interruption of traffic, the impact of which can be minimized by using fast-hardening and early-strength expansion joint materials. In this study, a fast-hardening polymer cement composite (PCC) was developed [...] Read more.
Bridges’ expansion joints are prone to damage during operation, and repairing them often requires interruption of traffic, the impact of which can be minimized by using fast-hardening and early-strength expansion joint materials. In this study, a fast-hardening polymer cement composite (PCC) was developed using sulfate aluminate cement and ordinary silicate cement as binding agents and polymer powder as admixture. To improve the crack resistance of the material, several types of fibers were added and the effects of different fiber types and admixtures on the crack resistance of the material were compared using SCB tests. The results showed that the best effect of improving the crack resistance of concrete was achieved with a volume fraction of 0.5% of basalt fibers. Then, a test method for the interfacial shear properties of PCC materials and ordinary concrete was established, and the cohesive force model was selected as the interface simulation parameter for finite element analysis and compared with experimental data to verify its feasibility. Finally, based on the previously obtained PCC material parameters, a solid model of the expansion joint anchorage zone was established to study the mechanical properties of the expansion joint anchorage zone with the application of fast-hardening PCC material. This research provides a new way to develop fast-hardening and early-strength expansion joint materials with high crack resistance. Full article
(This article belongs to the Section Building Materials, and Repair & Renovation)
Show Figures

Figure 1

Figure 1
<p>Standard sand gradation diagram.</p>
Full article ">Figure 2
<p>SCB specimens.</p>
Full article ">Figure 3
<p>SCB specimen fabrication.</p>
Full article ">Figure 4
<p>Interface compression shear test.</p>
Full article ">Figure 5
<p>Compressive strength at different ages.</p>
Full article ">Figure 6
<p>Effect of fiber content on compressive strength at different stages. (<span class="html-italic">S</span><sub>1</sub>: the degree of growth of <span class="html-italic">f<sub>c</sub></span> for this type of fiber concrete at 3–7 d; <span class="html-italic">S</span><sub>2</sub>: the degree of increase in <span class="html-italic">f<sub>c</sub></span> for this type of fiber concrete over 7–70 d).</p>
Full article ">Figure 7
<p>Failure of SCB specimen.</p>
Full article ">Figure 7 Cont.
<p>Failure of SCB specimen.</p>
Full article ">Figure 8
<p><span class="html-italic">P</span>-<span class="html-italic">u</span> curve of each specimen age.</p>
Full article ">Figure 9
<p>Cohesion–displacement fracture curve [<a href="#B36-buildings-13-02910" class="html-bibr">36</a>].</p>
Full article ">Figure 10
<p>Fracture energy analysis at different stages.</p>
Full article ">Figure 11
<p>Relationship between fracture toughness and age.</p>
Full article ">Figure 12
<p>Stiffness S diagram [<a href="#B35-buildings-13-02910" class="html-bibr">35</a>].</p>
Full article ">Figure 13
<p>Stiffness analysis of each period.</p>
Full article ">Figure 14
<p>Relationship between stiffness and age. (<span class="html-italic">S</span><sub>1</sub>: the degree of growth of <span class="html-italic">f<sub>c</sub></span> for BF fiber concrete at 8–70 d; <span class="html-italic">S</span><sub>2</sub>: the degree of increase in <span class="html-italic">f<sub>c</sub></span> for PAN fiber concrete over 8–70 d; <span class="html-italic">S</span><sub>3</sub>: the degree of increase in <span class="html-italic">f<sub>c</sub></span> for PP fiber concrete over 8–70 d).</p>
Full article ">Figure 15
<p>Interface compression–shear test result.</p>
Full article ">Figure 16
<p>Load-displacement curve.</p>
Full article ">Figure 17
<p>Compression–shear finite element model.</p>
Full article ">Figure 18
<p>Comparison of interface models in this paper.</p>
Full article ">Figure 19
<p>Cross-section of expansion joint installation.</p>
Full article ">Figure 20
<p>FEM of expansion joint.</p>
Full article ">Figure 21
<p>Unit division.</p>
Full article ">Figure 22
<p>Schematic diagram of the <span class="html-italic">D</span> value.</p>
Full article ">Figure 23
<p>Principal stress nephogram.</p>
Full article ">Figure 24
<p>The curve of maximum principal tensile stress in the anchorage zone versus <span class="html-italic">D</span> (<span class="html-italic">σ</span><sub>1<span class="html-italic">max</span></sub>-<span class="html-italic">D</span>).</p>
Full article ">Figure 25
<p>Waterfall diagram of maximum principal tensile stress in the anchorage zone versus <span class="html-italic">D</span>.</p>
Full article ">Figure 26
<p>The curve of maximum principal compressive stress in the anchorage zone versus <span class="html-italic">D</span> (<span class="html-italic">σ</span><sub>3<span class="html-italic">max</span></sub>-<span class="html-italic">D</span>).</p>
Full article ">Figure 27
<p>Principal compressive stress nephogram of anchorage zone (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 28
<p>Different materials’ curves of maximum principal stress in the anchorage zone versus <span class="html-italic">D</span> (<span class="html-italic">σ<sub>max</sub></span>-<span class="html-italic">D</span>).</p>
Full article ">
19 pages, 6951 KiB  
Article
Modified Lignin-Based Cement Solidifying Material for Improving Engineering Residual Soil
by Xiang Yu, Hongbo Lu, Jie Peng, Jinming Ren, Yongmin Wang and Junhao Chen
Materials 2023, 16(22), 7100; https://doi.org/10.3390/ma16227100 - 9 Nov 2023
Cited by 1 | Viewed by 828
Abstract
Although lignin improves the strength and modulus of soil, it is less active when unmodified, and it exhibits more limited effects on soils in combination with traditional Ca-based curing agents. Lignin-solidified soil also exhibits deficiencies, such as poor durability under dry–wet cycling conditions, [...] Read more.
Although lignin improves the strength and modulus of soil, it is less active when unmodified, and it exhibits more limited effects on soils in combination with traditional Ca-based curing agents. Lignin-solidified soil also exhibits deficiencies, such as poor durability under dry–wet cycling conditions, and thus, the amelioration effect is limited. This study investigated the enhancement of cement-solidified soil using hydroxylated lignin with sodium silicate and quicklime used as activators to improve the engineering performance and durability of the treated soil. Using respective cement, sodium silicate, quicklime, and lignin contents of 7%, 0.4%, 0.2%, and 0.2% with respect to the dry mass of the slag soil, the strength and cohesion of the composite-solidified soil were 1.5 times those of cement-solidified soil, whereas the internal friction angle increased by 5.1°. At a solidifying age of 14 d, the penetration resistance almost doubled, indicating a significant improvement in the bearing capacity of the soil. The results suggest that modified lignin-based admixtures may significantly enhance the performance of cement-solidified soil. The cement curing admixture used in this study provides theoretical and technological support for curing agent preparation and the utilization of slag. Full article
Show Figures

Figure 1

Figure 1
<p>Slag grading curve of the soil.</p>
Full article ">Figure 2
<p>Unconfined compressive strength at different material contents. (<b>a</b>) Different cement dosages and (<b>b</b>) 7% cement mixed with different materials.</p>
Full article ">Figure 3
<p>Relationship between shear strength and vertical stress.</p>
Full article ">Figure 4
<p>Relationships between the dry–wet cycling indices and number of cycles. (<b>a</b>) Cyclic and (<b>b</b>) cumulative mass losses and (<b>c</b>) unconfined compressive strength.</p>
Full article ">Figure 5
<p>Degrees of damage of the sample surfaces after ten dry–wet cycles: (<b>a</b>) 7% cement, (<b>b</b>) 7% cement + 0.2% quicklime, (<b>c</b>) 7% cement + 0.4% sodium silicate, or (<b>d</b>) composite solidifying agent.</p>
Full article ">Figure 6
<p>Variations in the shrinkage performance indices with time. (<b>a</b>) Shrinkage strain and (<b>b</b>) water loss rate.</p>
Full article ">Figure 7
<p>Schematics of the internal changes in the cement-based specimens. (<b>a</b>) Schematic of moisture migration in cured soil. (<b>b</b>) Schematic of the formation of negative pore pressure.</p>
Full article ">Figure 8
<p>Variations in the dynamic cone penetration indexes with age.</p>
Full article ">Figure 9
<p>Variations in penetration resistance (<span class="html-italic">Rs</span>): (<b>a</b>) 0 d hydraulic and (<b>b</b>) 0 d composite-consolidated soil <span class="html-italic">Rs</span> values, (<b>c</b>) 7 d hydraulic and (<b>d</b>) 7 d composite-consolidated soil <span class="html-italic">Rs</span> values, and (<b>e</b>) 14 d soil and (<b>f</b>) 14 d composite-consolidated soil <span class="html-italic">Rs</span> values.</p>
Full article ">Figure 10
<p>Results of X-ray diffraction analysis of the soils solidified using different materials: (<b>a</b>) plain soil and (<b>b</b>) plain soil and soils cured with different materials.</p>
Full article ">Figure 11
<p>Scanning electron microscopy micrographs of the soils solidified with different materials. (<b>a</b>) Plain soil, (<b>b</b>) 7% cement, (<b>c</b>) 7% cement + 0.2% quicklime, (<b>d</b>) 7% cement + 0.4% sodium silicate, and (<b>e</b>,<b>f</b>) composite solidifying agent.</p>
Full article ">Figure 12
<p>Infrared spectra of the soils solidified using different materials.</p>
Full article ">Figure 13
<p>Schematic of the mechanism of action of the solidifying material.</p>
Full article ">
393 KiB  
Proceeding Paper
The Formulation of Self-Compacting Concrete Mixtures Incorporating Diverse Cement Types
by Khandokar Md Rifat Hossain and Rupak Mutsuddy
Eng. Proc. 2023, 56(1), 41; https://doi.org/10.3390/ASEC2023-15238 - 26 Oct 2023
Viewed by 639
Abstract
Self-compacting concrete (SCC) is a highly flowable, self-leveling, and non-segregating type of concrete that requires no form of vibration to maintain its uniformity throughout the mixture as well as being able to perform in an outstanding manner in densely reinforced structures. The main [...] Read more.
Self-compacting concrete (SCC) is a highly flowable, self-leveling, and non-segregating type of concrete that requires no form of vibration to maintain its uniformity throughout the mixture as well as being able to perform in an outstanding manner in densely reinforced structures. The main objective of this study is to investigate the primary differences in the engineering properties of SCC using CEM-I, CEM-II/A-M, and CEM-II/B-M types of cement as the primary binding material. The properties of SCC, such as cohesiveness, stability, flowability, etc., can be modified by selecting definitive amounts of aggregates, cementitious materials, and viscosity-modifying admixtures. Therefore, it will highlight the effects of the mechanical and flow properties of the concrete mix due to the change in cement type with a similar composition and volumetric ratio to other constituent materials. The flow properties were validated using the V-funnel test, L-box test, T-500 test, and slump flow test. A comparative result highlighting the strength response, i.e., the compressive, tensile, and flexural strength of the mix designs, was recorded at 28 days, and correlations among these values were established and analyzed. Full article
(This article belongs to the Proceedings of The 4th International Electronic Conference on Applied Sciences)
Show Figures

Figure 1

Figure 1
<p>Load vs. deflection curve for flexural strength.</p>
Full article ">
13 pages, 3277 KiB  
Article
Determination of the Basic Geotechnical Parameters of Blast-Furnace Slag from the Kremnica Region
by Roman Bulko, Soňa Masarovičová and Filip Gago
Materials 2023, 16(17), 5966; https://doi.org/10.3390/ma16175966 - 31 Aug 2023
Cited by 2 | Viewed by 987
Abstract
A decisive aspect of site evaluation for construction is the presence of anthropogenic materials occurring in the geological environment. The geotechnical properties of blast-furnace slag were investigated as a potential substitute for aggregates in the construction industry. The basic geotechnical parameters of the [...] Read more.
A decisive aspect of site evaluation for construction is the presence of anthropogenic materials occurring in the geological environment. The geotechnical properties of blast-furnace slag were investigated as a potential substitute for aggregates in the construction industry. The basic geotechnical parameters of the slag were determined, which are critical for evaluating its stability, environmental impact, and usability in geotechnical construction. The research focused on monitoring the physical and mechanical properties of the two samples, and also included mineralogical analysis. The obtained results demonstrated that the slag belongs to the category of poorly graded gravel, G2/GP, and gravel with an admixture of fine-grained soil, G3/G-F. In addition, other important parameters, such as the water disintegration of the slag aggregate, the minimum and maximum bulk densities, the California bearing ratio (CBR), the oedometric modulus (Eoed), and shear tests (the angle of internal friction φ and cohesion c), were determined. The results from this paper provide important information for the proper management of blast-furnace slag so to minimize its environmental impact and achieve sustainability in the mining industry. At the same time, it enables a better understanding of the use of slag as a substitute for aggregates in geotechnical tasks. Despite its local importance in relation to the investigated case, the presented study has significant educational and scientific value for the construction sector, where it is necessary to evaluate anthropogenic activities and materials. Full article
Show Figures

Figure 1

Figure 1
<p>Porous structure of slag.</p>
Full article ">Figure 2
<p>Diffractogram of a slag sample (Fa—fayalite, Hd—hedenbergite, Ge—gehlenite, Tep—tephroite, Wo—wollastonite, Sil—sillimanite, Ttn—titanite, Anl—analcin) [<a href="#B34-materials-16-05966" class="html-bibr">34</a>].</p>
Full article ">Figure 3
<p>Grading curve of slag under 32 mm.</p>
Full article ">Figure 4
<p>Penetration cylinder impressions inside the red circle—<span class="html-italic">CBR</span> test: <span class="html-italic">G</span>3<span class="html-italic">/G-F</span> is on the left and <span class="html-italic">G</span>2<span class="html-italic">/GP</span> is on the right.</p>
Full article ">Figure 5
<p>Dependence of the design modulus of elasticity <span class="html-italic">E<sub>p,n</sub></span> on the <span class="html-italic">CBR</span> strength according to [<a href="#B38-materials-16-05966" class="html-bibr">38</a>].</p>
Full article ">Figure 6
<p>Determination of oedometric modulus <span class="html-italic">E<sub>oed</sub></span>.</p>
Full article ">Figure 7
<p>Evaluation of shear box tests of blast-furnace slag. (<b>a</b>) <span class="html-italic">G</span>3<span class="html-italic">/G-F</span>; (<b>b</b>) <span class="html-italic">G</span>2<span class="html-italic">/GP</span>.</p>
Full article ">Figure 7 Cont.
<p>Evaluation of shear box tests of blast-furnace slag. (<b>a</b>) <span class="html-italic">G</span>3<span class="html-italic">/G-F</span>; (<b>b</b>) <span class="html-italic">G</span>2<span class="html-italic">/GP</span>.</p>
Full article ">
18 pages, 9293 KiB  
Article
Curing Mechanisms of Polymeric Nano-Copolymer Subgrade
by Shuang Shi, Miao Wang, Linhao Gu, Xiang Chen and Yanning Zhang
Materials 2023, 16(12), 4316; https://doi.org/10.3390/ma16124316 - 11 Jun 2023
Viewed by 1221
Abstract
The mechanical properties of the subgrade have a significant impact on the service life and pavement performance of the superstructure of pavement. By adding admixtures and via other means to strengthen the adhesion between soil particles, the strength and stiffness of the soil [...] Read more.
The mechanical properties of the subgrade have a significant impact on the service life and pavement performance of the superstructure of pavement. By adding admixtures and via other means to strengthen the adhesion between soil particles, the strength and stiffness of the soil can be improved to ensure the long-term stability of pavement structures. In this study, a mixture of polymer particles and nanomaterials was used as a curing agent to examine the curing mechanism and mechanical properties of subgrade soil. Using microscopic experiments, the strengthening mechanism of solidified soil was analyzed with scanning electron microscopy (SEM), energy-dispersive spectroscopy (EDS), Fourier infrared spectroscopy (FTIR), and X-ray diffraction (XDR). The results showed that with the addition of the curing agent, small cementing substances on the surface of soil minerals filled the pores between minerals. At the same time, with an increase in the curing age, the colloidal particles in the soil increased, and some of them formed large aggregate structures that gradually covered the surface of the soil particles and minerals. By enhancing the cohesiveness and integrity between different particles, the overall structure of the soil became denser. Through pH tests, it was found that the age had a certain effect on the pH of solidified soil, but the effect was not obvious. Through the comparative analysis of elements in plain soil and solidified soil, it was found that no new chemical elements were produced in the solidified soil, indicating that the curing agent does not have negative impacts on the environment. Full article
(This article belongs to the Section Advanced Composites)
Show Figures

Figure 1

Figure 1
<p>Dry density vs. moisture content.</p>
Full article ">Figure 2
<p>Curing agent.</p>
Full article ">Figure 3
<p>SEM sample preparation: (<b>a</b>) specimen preparation; (<b>b</b>) sample after vacuum and gold plating.</p>
Full article ">Figure 4
<p>pH examination.</p>
Full article ">Figure 5
<p>Fourier infrared spectroscopy (FTIR) test. (<b>a</b>) Potassium bromide mixed grinding. (<b>b</b>) KBr pellet. (<b>c</b>) ATR.</p>
Full article ">Figure 6
<p>Smart X-ray diffractometer with sample preparation. (<b>a</b>) Intelligent X-ray diffractometer. (<b>b</b>) Observed specimen.</p>
Full article ">Figure 7
<p>Scanning electron microscope test results. (<b>a</b>) Microstructure of plain soil. (<b>b</b>) Microstructure of solidified soil with 0.015% was preserved for 7 days. (<b>c</b>) Microstructure of solidified soil with 0.025% curing agent dosage was preserved for 7 days. (<b>d</b>) Microstructure of solidified soil with 0.025% curing agent dosage was preserved for 21 days.</p>
Full article ">Figure 7 Cont.
<p>Scanning electron microscope test results. (<b>a</b>) Microstructure of plain soil. (<b>b</b>) Microstructure of solidified soil with 0.015% was preserved for 7 days. (<b>c</b>) Microstructure of solidified soil with 0.025% curing agent dosage was preserved for 7 days. (<b>d</b>) Microstructure of solidified soil with 0.025% curing agent dosage was preserved for 21 days.</p>
Full article ">Figure 7 Cont.
<p>Scanning electron microscope test results. (<b>a</b>) Microstructure of plain soil. (<b>b</b>) Microstructure of solidified soil with 0.015% was preserved for 7 days. (<b>c</b>) Microstructure of solidified soil with 0.025% curing agent dosage was preserved for 7 days. (<b>d</b>) Microstructure of solidified soil with 0.025% curing agent dosage was preserved for 21 days.</p>
Full article ">Figure 7 Cont.
<p>Scanning electron microscope test results. (<b>a</b>) Microstructure of plain soil. (<b>b</b>) Microstructure of solidified soil with 0.015% was preserved for 7 days. (<b>c</b>) Microstructure of solidified soil with 0.025% curing agent dosage was preserved for 7 days. (<b>d</b>) Microstructure of solidified soil with 0.025% curing agent dosage was preserved for 21 days.</p>
Full article ">Figure 8
<p>Variation in pH of solidified soil. (<b>a</b>) Variation in pH of solidified soil with age. (<b>b</b>) Variation in pH of cement-cured soil with age.</p>
Full article ">Figure 9
<p>Spectral analysis of plain soil. (<b>a</b>) Microstructure of plain soil (10,000 times magnification). (<b>b</b>) Elemental analysis of plain soil.</p>
Full article ">Figure 10
<p>Spectral analysis of solidified soil. (<b>a</b>) Microstructure of solidified soil (10,000 times magnification). (<b>b</b>) Elemental analysis of solidified soil.</p>
Full article ">Figure 11
<p>Functional group analysis of curing agent.</p>
Full article ">Figure 12
<p>Functional group analysis of plain soil.</p>
Full article ">Figure 13
<p>Functional group analysis of solidified soil.</p>
Full article ">Figure 14
<p>X-ray diffraction analysis results of (<b>a</b>) plain and (<b>b</b>) solidified soil.</p>
Full article ">
22 pages, 10610 KiB  
Article
Numerical Simulation of Fatigue Life of Rubber Concrete on the Mesoscale
by Xianfeng Pei, Xiaoyu Huang, Houmin Li, Zhou Cao, Zijiang Yang, Dingyi Hao, Kai Min, Wenchao Li, Cai Liu, Shuai Wang and Keyang Wu
Polymers 2023, 15(9), 2048; https://doi.org/10.3390/polym15092048 - 25 Apr 2023
Cited by 6 | Viewed by 1583
Abstract
Rubber concrete (RC) exhibits high durability due to the rubber admixture. It is widely used in a large number of fatigue-resistant structures. Mesoscale studies are used to study the composition of polymers, but there is no method for fatigue simulation of RC. Therefore, [...] Read more.
Rubber concrete (RC) exhibits high durability due to the rubber admixture. It is widely used in a large number of fatigue-resistant structures. Mesoscale studies are used to study the composition of polymers, but there is no method for fatigue simulation of RC. Therefore, this paper presents a finite element modeling approach to study the fatigue problem of RC on the mesoscale, which includes the random generation of the main components of the RC mesoscale structure. We also model the interfacial transition zone (ITZ) of aggregate mortar and the ITZ of rubber mortar. This paper combines the theory of concrete damage to plastic with the method of zero-thickness cohesive elements in the ITZ, and it is a new numerical approach. The results show that the model can simulate reasonably well the random damage pattern after RC beam load damage. The damage occurred in the middle of the beam span and tended to follow the ITZ. The model can predict the fatigue life of RC under various loads. Full article
(This article belongs to the Special Issue Modeling and Simulation of Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>Flowchart of aggregate generation.</p>
Full article ">Figure 2
<p>The uniaxial stress–strain response: (<b>a</b>) tensile; (<b>b</b>) compressive.</p>
Full article ">Figure 3
<p>Redefinition of tensile stiffness after damage.</p>
Full article ">Figure 4
<p>Three types of CE: (<b>a</b>) shared nodes; (<b>b</b>) tie constraint; (<b>c</b>) contact interaction.</p>
Full article ">Figure 5
<p>Aggregates and ITZ generated in ABAQUS.</p>
Full article ">Figure 6
<p>Traction separation relationship.</p>
Full article ">Figure 7
<p>RC beams with different dosing: (<b>a</b>) 0; (<b>b</b>) 2.5%; (<b>c</b>) 5%; (<b>d</b>) 7.5%; (<b>e</b>) 10%.</p>
Full article ">Figure 8
<p>Load loading schematic.</p>
Full article ">Figure 9
<p>Experimental and simulated fatigue life.</p>
Full article ">Figure 10
<p>Load–deflection curve.</p>
Full article ">Figure 11
<p>Damage to ordinary concrete and location of damage sampling.</p>
Full article ">Figure 12
<p>RC SDEG with different mixes for static pressure loading: (<b>a</b>) 0; (<b>b</b>) 2.5%; (<b>c</b>) 5%; (<b>d</b>) 7.5%; (<b>e</b>) 10%.</p>
Full article ">Figure 12 Cont.
<p>RC SDEG with different mixes for static pressure loading: (<b>a</b>) 0; (<b>b</b>) 2.5%; (<b>c</b>) 5%; (<b>d</b>) 7.5%; (<b>e</b>) 10%.</p>
Full article ">Figure 13
<p>SDEG of ordinary concrete under different stress levels: (<b>a</b>) S = 0.9; (<b>b</b>) S = 0.85; (<b>c</b>) S = 0.8; (<b>d</b>) S = 0.75.</p>
Full article ">Figure 14
<p>SDEG of RC-2.5 under different stress levels: (<b>a</b>) S = 0.9; (<b>b</b>) S = 0.85; (<b>c</b>) S = 0.8; (<b>d</b>) S = 0.75.</p>
Full article ">Figure 15
<p>SDEG of RC-5 under different stress levels: (<b>a</b>) S = 0.9; (<b>b</b>) S = 0.85; (<b>c</b>) S = 0.8; (<b>d</b>) S = 0.75.</p>
Full article ">Figure 16
<p>SDEG of RC-7.5 under different stress levels: (<b>a</b>) S = 0.9; (<b>b</b>) S = 0.85; (<b>c</b>) S = 0.8; (<b>d</b>) S = 0.75.</p>
Full article ">Figure 17
<p>SDEG of RC-10 under different stress levels: (<b>a</b>) S = 0.9; (<b>b</b>) S = 0.85; (<b>c</b>) S = 0.8; (<b>d</b>) S = 0.75.</p>
Full article ">Figure 18
<p>Fatigue life of different rubber doping at different stress levels.</p>
Full article ">Figure 19
<p>Relationship between stress level and fatigue life of different rubber doping.</p>
Full article ">Figure 20
<p>Fitting relationship between different rubber doping stress levels and fatigue life.</p>
Full article ">
22 pages, 9079 KiB  
Article
Sand Transport with Cohesive Admixtures…—Laboratory Tests and Modeling
by Jerzy Zawisza, Iwona Radosz, Jarosław Biegowski and Leszek M. Kaczmarek
Water 2023, 15(4), 804; https://doi.org/10.3390/w15040804 - 18 Feb 2023
Cited by 2 | Viewed by 2141
Abstract
The paper presents results of experimental and theoretical studies on transport of water-sand mixtures in steady flow with small amounts of cohesive fractions. The experiments were carried out for sand alone and with cohesive admixtures in the form of clay in the amount [...] Read more.
The paper presents results of experimental and theoretical studies on transport of water-sand mixtures in steady flow with small amounts of cohesive fractions. The experiments were carried out for sand alone and with cohesive admixtures in the form of clay in the amount of 5, 10, 15 and 20% by weight. The amount of sand fractions retained in the trap and along the control area was measured. The experimental results were compared with the calculation results for transport rate of sand fractions. An intended model of the vertical structure of both sand velocity and concentration as well as vertical mixing and sorting is proposed here in order to determine the influence of cohesive admixtures on the transport of sand fractions. Hence the reduction of sand fractions transport due to cohesion forces is included. The agreement of sand transport calculations according to the extended model with measured results and experimental data from literature was achieved within plus/minus a factor of 2. Full article
(This article belongs to the Special Issue Sediment Transport, Budgets and Quality in Riverine Environments)
Show Figures

Figure 1

Figure 1
<p>Experimental setup for steady flow measurements Gdańsk 2021.</p>
Full article ">Figure 2
<p>Grain size distributions used in experiments.</p>
Full article ">Figure 3
<p>Results of transport rate measurements of sand fractions of sediments with cohesive admixtures obtained in Gdańsk 2021 experiment with the approximations of mean values of repeated tests by curves with a coefficients of determination.</p>
Full article ">Figure 4
<p>Transport of sand fractions in the experiments by De Sutter et al. [<a href="#B48-water-15-00804" class="html-bibr">48</a>].</p>
Full article ">Figure 5
<p>Vertical structure of: (<b>a</b>) sediment transport profile with velocity and concentration of the <span class="html-italic">i</span>-th fraction of sediment; (<b>b</b>) shear stress profile with cohesion stress <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mrow> <mi>c</mi> <mi>o</mi> <mi>h</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Flow charts of numerical algorithms for calculations of sediment transport and the mobile-bed effect parameter <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mn>0</mn> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Influence of cohesive forces on calculations of the parameter <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mn>0</mn> <mi>c</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math> depending on the content of clay in the sandy deposit data: <math display="inline"><semantics> <mrow> <msub> <msup> <mi>u</mi> <mo>′</mo> </msup> <mrow> <mi>f</mi> <mo>∗</mo> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>u</mi> <msub> <mo>′</mo> <mrow> <mi>f</mi> <mo>∗</mo> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mi>r</mi> </msub> <mo>=</mo> <msub> <mi>d</mi> <mrow> <mn>50</mn> </mrow> </msub> <mo>=</mo> <mn>0.32</mn> </mrow> </semantics></math> mm from the experiment of De Sutter et al. [<a href="#B48-water-15-00804" class="html-bibr">48</a>].</p>
Full article ">Figure 8
<p>Influence of cohesive forces of calculations of velocity profiles in: (<b>a</b>) contact layer; (<b>b</b>) in dense layer and concentration profiles in: (<b>c</b>) contact layer, (<b>d</b>) in dense layer; data: <math display="inline"><semantics> <mrow> <mi>u</mi> <msub> <mo>′</mo> <mrow> <mi>f</mi> <mo>∗</mo> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>u</mi> <msub> <mo>′</mo> <mrow> <mi>f</mi> <mo>∗</mo> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> from the experiment of De Sutter et al. [<a href="#B48-water-15-00804" class="html-bibr">48</a>].</p>
Full article ">Figure 9
<p>Influence of disregarding cohesion on transport rate calculations for De Sutter [<a href="#B48-water-15-00804" class="html-bibr">48</a>].</p>
Full article ">Figure 10
<p>Comparison of sand transport calculations with Gdańsk 2021 measurements.</p>
Full article ">Figure 11
<p>Comparison of sand transport calculations with Gdańsk 2021 measurements the approximation of mean values of repeated tests by linear curve with a coefficient of determination <math display="inline"><semantics> <mrow> <msup> <mi>R</mi> <mn>2</mn> </msup> <mo>=</mo> <mn>0.9692</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 12
<p>Comparison of sand transport calculations with measurements by De Sutter et al. [<a href="#B48-water-15-00804" class="html-bibr">48</a>] with the approximation of mean values of repeated tests by linear curve with a coefficient of determination <math display="inline"><semantics> <mrow> <msup> <mi>R</mi> <mn>2</mn> </msup> <mo>=</mo> <mn>0.9145</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 13
<p>Comparison of sand transport calculations with measurements by Torfs [<a href="#B63-water-15-00804" class="html-bibr">63</a>] the approximation of mean values of repeated tests by linear curve with a coefficient of determination <math display="inline"><semantics> <mrow> <msup> <mi>R</mi> <mn>2</mn> </msup> <mo>=</mo> <mn>0.9873</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 14
<p>Comparison of sand transport calculations with measurements by Alvarez -Hernandez [<a href="#B45-water-15-00804" class="html-bibr">45</a>] the approximation of mean values of repeated tests by linear curve with a coefficient of determination <math display="inline"><semantics> <mrow> <msup> <mi>R</mi> <mn>2</mn> </msup> <mo>=</mo> <mn>0.9006</mn> </mrow> </semantics></math>. Addition of 20% clay with two densities c = 24 g/L and 30 g/L.</p>
Full article ">
13 pages, 1241 KiB  
Article
The Application of Recycled Epoxy Plastic Sheets Waste to Replace Concrete in Urban Construction and Building
by Bowen Qi, Shouwu Gao and Peilong Xu
Processes 2023, 11(1), 201; https://doi.org/10.3390/pr11010201 - 8 Jan 2023
Cited by 25 | Viewed by 2260
Abstract
Epoxy plastic, a form of epoxy resin, is widely employed in a variety of sectors due to its superior mechanical qualities and adaptability. The application of waste epoxy plastic in urban highway construction has been a major topic of study. To conduct this [...] Read more.
Epoxy plastic, a form of epoxy resin, is widely employed in a variety of sectors due to its superior mechanical qualities and adaptability. The application of waste epoxy plastic in urban highway construction has been a major topic of study. To conduct this study, epoxy polymers are mixed with concrete to enhance the thermal and compressive resistance and tensile strength, which acts as a substitute for conventional cements. The experimental results indicate that ER concrete has good cohesive qualities since it does not collapse or peel, and the nature of the epoxy plastic guarantees that ER concrete has great mechanical capabilities due to the strong bond between the epoxy resin and the fibres. In terms of frost resistance, granular concrete with a 10% ER additive has a mass loss rate between 0.3% and 0.12% and a strength loss rate between 3.55 and 9.4%, outperforming conventional concrete. When often loaded by traffic, ER concrete exhibits no substantial permanent deformation, and its fatigue damage rate is superior to that of ordinary concrete. In total, 10% admixture of ER concrete may efficiently fulfil BPN (British Pendulum Number) and structural depth standards, while greatly improving the road’s skid resistance. In addition, its modulus of elasticity, deformation capacity, and high-temperature stability are superior to those of conventional concrete. Full article
(This article belongs to the Section Sustainable Processes)
Show Figures

Figure 1

Figure 1
<p>Crack fracture mode. (<b>a</b>) tensile (<b>b</b>) sliding (<b>c</b>) tearing.</p>
Full article ">Figure 2
<p>Comparison curve of slump with different dosage and particle size.</p>
Full article ">Figure 3
<p>Splitting tensile strength of ER concrete.</p>
Full article ">Figure 4
<p>Comparison diagram of fatigue damage rate of concrete and rice husk concrete in this study.</p>
Full article ">Figure 5
<p>Slip resistance of ER concrete.</p>
Full article ">Figure 6
<p>Dynamic stability of concrete with different ER.</p>
Full article ">
17 pages, 7197 KiB  
Article
Multi-Scale Characterization of High-Temperature Properties and Thermal Storage Stability Performance of Discarded-Mask-Modified Asphalt
by Yuanle Li, Bing Hui, Xinyi Yang, Huimin Wang, Ning Xu, Ponan Feng, Ziye Ma and Hainian Wang
Materials 2022, 15(21), 7593; https://doi.org/10.3390/ma15217593 - 28 Oct 2022
Cited by 4 | Viewed by 1551
Abstract
In the context of the global pandemic of COVID-19, the use and disposal of medical masks have created a series of ethical and environmental issues. The purpose of this paper is to study and evaluate the high temperature properties and thermal storage stability [...] Read more.
In the context of the global pandemic of COVID-19, the use and disposal of medical masks have created a series of ethical and environmental issues. The purpose of this paper is to study and evaluate the high temperature properties and thermal storage stability of discarded-mask (DM)-modified asphalt from a multi-scale perspective using molecular dynamics (MD) simulation and experimental methods. A series of tests was conducted to evaluate the physical, rheological, thermal storage stability and microscopic properties of the samples. These tests include softening point, rotational viscosity, dynamic shear rheology (DSR), Fourier transform infrared (FT-IR) spectroscopy and molecular dynamics simulation. The results showed that the DM modifier could improve the softening point, rotational viscosity and rutting factor of the asphalt. After thermal storage, the DM-modified asphalt produced segregation. The difference in the softening point between the top and bottom of the sample increased from 2.2 °C to 17.1 °C when the DM modifier admixture was increased from 1% to 4%. FT-IR test results showed that the main component of the DM modifier was polypropylene, and the DM-modified asphalt was mainly a physical co-blending process. MD simulation results show that the DM modifier can increase the cohesive energy density (CED) and reduce the fractional free volume (FFV) of asphalt and reduce the binding energy between base asphalt and DM modifier. Multi-scale characterization reveals that DM modifiers can improve the high temperature performance and reduce the thermal storage stability of asphalt. It is noteworthy that both macroscopic tests and microscopic simulations show that 1% is an acceptable dosage level. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>DM-modified asphalt preparation process.</p>
Full article ">Figure 2
<p>Flowchart of the experimental plan.</p>
Full article ">Figure 3
<p>Molecular structure of base asphalt.</p>
Full article ">Figure 4
<p>Physical properties of DM-modified asphalt. (<b>a</b>) Softening point and (<b>b</b>) rotational viscosity (135 °C, 155 °C and 175 °C).</p>
Full article ">Figure 5
<p>Temperature scanning test results. (<b>a</b>) Phase angle; (<b>b</b>) complex modulus; (<b>c</b>) rutting factor.</p>
Full article ">Figure 6
<p>Softening point of DM-modified asphalt after thermal storage.</p>
Full article ">Figure 7
<p>DM-modified asphalt VS index comparison results.</p>
Full article ">Figure 8
<p>DM-modified asphalt RS index result.</p>
Full article ">Figure 9
<p>FTIR spectra of discarded medical masks.</p>
Full article ">Figure 10
<p>FTIR spectra of DM-modified asphalt.</p>
Full article ">Figure 11
<p>Solubility parameters of polypropylene.</p>
Full article ">Figure 12
<p>Steady-state DM-modified asphalt model. (<b>a</b>) DM1-modified asphalt model; (<b>b</b>) DM2-modified asphalt model; (<b>c</b>) DM3-modified asphalt model; (<b>d</b>) DM4-modified asphalt model. The blue part is the polypropylene chain.</p>
Full article ">Figure 12 Cont.
<p>Steady-state DM-modified asphalt model. (<b>a</b>) DM1-modified asphalt model; (<b>b</b>) DM2-modified asphalt model; (<b>c</b>) DM3-modified asphalt model; (<b>d</b>) DM4-modified asphalt model. The blue part is the polypropylene chain.</p>
Full article ">Figure 13
<p>Model reasonableness verification. (<b>a</b>) Energy state of the model optimization process and (<b>b</b>) radial distribution function (RDF).</p>
Full article ">Figure 14
<p>The cohesive energy density of the model.</p>
Full article ">Figure 15
<p>Morphology of the free volume at 331 K. (<b>a</b>) Schematic of the free volume calculation method; (<b>b</b>) base asphalt.</p>
Full article ">
21 pages, 33511 KiB  
Article
A Study on the Relationships between Water Film Thickness, Fresh Properties, and Mechanical Properties of Cement Paste Containing Superfine Basalt Powder (SB)
by Hengrui Liu, Zhenghong Tian and Haoyue Fan
Materials 2021, 14(24), 7592; https://doi.org/10.3390/ma14247592 - 10 Dec 2021
Cited by 2 | Viewed by 1456
Abstract
In this paper, the effect of a newly developed superfine basalt powder (SB) on the fresh and mechanical properties of cement paste was studied. The concept of water film thickness (WFT) was cited to explain the influence of SB on fresh and mechanical [...] Read more.
In this paper, the effect of a newly developed superfine basalt powder (SB) on the fresh and mechanical properties of cement paste was studied. The concept of water film thickness (WFT) was cited to explain the influence of SB on fresh and mechanical properties and related mathematical model formulas were established. In addition, the relationship between the fresh properties and mechanical properties of paste was also explored. The results indicated that SB can improve the segregation resistance and cohesiveness. The maximum improvement rate relative to the control cement paste was 75.4% and 50.4%, respectively. The 5% SB and 10% SB reduced the fluidity in the range of 4.1–68.7% but increased the early and late compressive strength in the range of 1.2–25.7% compared to control cement paste under different water/cementitious materials (W/CM) ratios. However, the influence of 20% SB on fluidity and compressive strength was opposite to the above behavior, and the increase rate and decrease rate were 1.8–11.8% and 1.1–13.9% respectively. The WFT was the most important factor that determined the compressive strength, rheological parameters, and flow parameters of paste containing SB, while the substitute content of SB and WFT together determined the bleeding rate and cohesiveness. Among them, the correlation between bleeding rate and WFT increased with time. The empirical mathematical models between WFT, fresh properties, and compressive strength were established and verified by other mineral admixtures, which were successfully extended and applied to the entire field of cement-based materials Full article
Show Figures

Figure 1

Figure 1
<p>Particle size distributions of OPC and SB.</p>
Full article ">Figure 2
<p>SEM morphology of SB.</p>
Full article ">Figure 3
<p>Variations of the void ratio and packing density at different SB substitutions.</p>
Full article ">Figure 4
<p>Variations of the WFT with W/CM ratio at different SB substitutions.</p>
Full article ">Figure 5
<p>Flow rate and flow spread versus W/CM ratio (<b>a</b>,<b>b</b>) and WFT (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 6
<p>Yield stress and apparent viscosity versus W/CM ratio (<b>a</b>,<b>b</b>) and WFT (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 7
<p>Cohesiveness versus W/CM ratio (<b>a</b>) and WFT (<b>b</b>).</p>
Full article ">Figure 8
<p>Bleeding rate versus W/CM ratio (<b>a</b>,<b>b</b>) and WFT (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 9
<p>Compressive strength versus W/CM ratio (<b>a</b>) and WFT (<b>b</b>–<b>d</b>).</p>
Full article ">Figure 10
<p>Flow rate and flow spread versus apparent viscosity (<b>c</b>,<b>d</b>) and yield stress (<b>a</b>,<b>b</b>).</p>
Full article ">Figure 11
<p>Flow rate and flow spread versus 1 day (<b>a</b>,<b>b</b>), 7 days (<b>c</b>,<b>d</b>), and 28 days (<b>e</b>,<b>f</b>) of compressive strength.</p>
Full article ">Figure 12
<p>Apparent viscosity and yield stress versus 1 day (<b>a</b>,<b>b</b>), 7 days (<b>c</b>,<b>d</b>), and 28 days (<b>e</b>,<b>f</b>) of compressive strength.</p>
Full article ">Figure 13
<p>Flow rate and flow spread versus 1 h (<b>a,b</b>) and 3 h (<b>c,d</b>) bleeding rate.</p>
Full article ">Figure 14
<p>Apparent viscosity and yield stress versus 1 h (<b>a,b</b>) and 3 h (<b>c,d</b>) bleeding rate.</p>
Full article ">Figure 15
<p>Flow rate (<b>a</b>), flow spread (<b>b</b>), apparent viscosity (<b>c</b>), and yield stress (<b>d</b>) versus cohesiveness.</p>
Full article ">Figure 16
<p>Bleeding rate and cohesiveness versus 1 day (<b>a</b>,<b>b</b>), 7 days (<b>c</b>,<b>d</b>), and 28 days (<b>e</b>,<b>f</b>) of compressive strength.</p>
Full article ">Figure 17
<p>SEM photography and particle size distribution of silica powder.</p>
Full article ">Figure 18
<p>Mathematical empirical model verification (<b>a</b>,<b>b</b>): (<b>a</b>) The mathematical empirical models between WFT and fresh properties; (<b>b</b>) The mathematical empirical models between WFT, fresh properties and compressive strength.</p>
Full article ">Figure 18 Cont.
<p>Mathematical empirical model verification (<b>a</b>,<b>b</b>): (<b>a</b>) The mathematical empirical models between WFT and fresh properties; (<b>b</b>) The mathematical empirical models between WFT, fresh properties and compressive strength.</p>
Full article ">
Back to TopTop